166
University of Massachusetts Amherst University of Massachusetts Amherst ScholarWorks@UMass Amherst ScholarWorks@UMass Amherst Open Access Dissertations 5-2010 Topics in Multivariate Time Series Analysis: Statistical Control, Topics in Multivariate Time Series Analysis: Statistical Control, Dimension Reduction Visualization and Their Business Dimension Reduction Visualization and Their Business Applications Applications Xuan Huang University of Massachusetts Amherst Follow this and additional works at: https://scholarworks.umass.edu/open_access_dissertations Part of the Management Sciences and Quantitative Methods Commons Recommended Citation Recommended Citation Huang, Xuan, "Topics in Multivariate Time Series Analysis: Statistical Control, Dimension Reduction Visualization and Their Business Applications" (2010). Open Access Dissertations. 204. https://scholarworks.umass.edu/open_access_dissertations/204 This Open Access Dissertation is brought to you for free and open access by ScholarWorks@UMass Amherst. It has been accepted for inclusion in Open Access Dissertations by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact [email protected].

Topics in Multivariate Time Series Analysis: Statistical

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Topics in Multivariate Time Series Analysis: Statistical

University of Massachusetts Amherst University of Massachusetts Amherst

ScholarWorks@UMass Amherst ScholarWorks@UMass Amherst

Open Access Dissertations

5-2010

Topics in Multivariate Time Series Analysis: Statistical Control, Topics in Multivariate Time Series Analysis: Statistical Control,

Dimension Reduction Visualization and Their Business Dimension Reduction Visualization and Their Business

Applications Applications

Xuan Huang University of Massachusetts Amherst

Follow this and additional works at: https://scholarworks.umass.edu/open_access_dissertations

Part of the Management Sciences and Quantitative Methods Commons

Recommended Citation Recommended Citation Huang, Xuan, "Topics in Multivariate Time Series Analysis: Statistical Control, Dimension Reduction Visualization and Their Business Applications" (2010). Open Access Dissertations. 204. https://scholarworks.umass.edu/open_access_dissertations/204

This Open Access Dissertation is brought to you for free and open access by ScholarWorks@UMass Amherst. It has been accepted for inclusion in Open Access Dissertations by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact [email protected].

Page 2: Topics in Multivariate Time Series Analysis: Statistical

TOPICS IN MULTIVARIATE TIME SERIES ANALYSIS: STATISTICAL CONTROL, DIMENSION REDUCTION, VISUALIZATION

AND THEIR BUSINESS APPLICATIONS

A Dissertation Presented

by

XUAN HUANG

Submitted to the Graduate School of the University of Massachusetts Amherst in partial fulfillment

of the requirements for the degree of

DOCTOR OF PHILOSOPHY

May 2010

Management Management Science

Page 3: Topics in Multivariate Time Series Analysis: Statistical

© Copyright by Xuan Huang 2010

All Rights Reserved

Page 4: Topics in Multivariate Time Series Analysis: Statistical

TOPICS IN MULTIVARIATE TIME SERIES ANALYSIS: STATISTICAL CONTROL, DIMENSION REDUCTION, VISUALIZATION

AND THEIR BUSINESS APPLICATIONS

A Dissertation Presented

by

XUAN HUANG

Approved as to style and content by: _______________________________________ Iqbal Agha, Chair _______________________________________ John Buonaccorsi, Member _______________________________________ Robert Nakosteen, Member

_________________________________________ Iqbal Agha, Department Chair Department of Finance & Operations Management

Page 5: Topics in Multivariate Time Series Analysis: Statistical

DEDICATION

In memory of my doctoral advisor

Søren Bisgaard

Page 6: Topics in Multivariate Time Series Analysis: Statistical

v

ACKNOWLEDGEMENTS

First and foremost, I have to thank Professor Søren Bisgaard, whose close

guidance accompanied every single step of my progress. Without him the completion of

this thesis wouldn’t have been possible.

Professor Bisgaard had always been there – my first conference presentation, my

first journal acceptance, qualification exams, and dissertation proposal – to steer me in

the correct direction or simply cheer me on. Unfortunately he won’t see the finished

version of my thesis, but I can almost picture it: he would give it another careful and

thorough proofreading, for the, say, fifteenth time. He would find a misplaced

punctuation, a misused article or a misaligned formula, and he would point them out in

the most gentle and kind way. For the papers we coauthored, we always iterated

proofreading – our record was “version 17,” and we had the “final of final version.” He

always taught me to be careful, accurate, and precise. With this thought in mind, I

endlessly proofread, in fear that the version would not meet his standard.

I am grateful for what he taught me: to never lose sight of real world applications,

to not easily give up ideas in doubt of how other researchers would receive them, and that

research is a passion rather than a profession. I am grateful that he respected me as a

friend and placed my interest before my obligation as his research assistant. He funded

me to finish an MSc in Statistics. He encouraged me to learn beyond our research scope.

And there were many dinners, barbeques, and afternoon teas at his house. We talked

about economics, science, technologies, as well as music, jokes, and trips to Disney

World.

Page 7: Topics in Multivariate Time Series Analysis: Statistical

vi

And these fun memories wouldn’t have been so sweet without Sue Ellen, his wife.

She welcomes and loves us graduate students as family members; she prepares wonderful

food and provides warm conversations for our every visit. Her unbreakable strength, in

accompanying Professor Bisgaard battling with cancer, is admirable and inspiring.

Special thanks also go to my dissertation committee. Thanks to Professor Iqbal

Agha, who, in his sabbatical, stepped up and shouldered the responsibility of being the

chair of my committee soon after Professor Bisgaard’s passing. Thanks to Professor John

Buonaccorsi, whose teaching shaped my understanding of statistics and inspired my

interest in statistical research. Thanks to Professor Robert Nakosteen, who has always

been so gracious to take time out of his busy schedule and flexible with my often short-

notice requests. Of course, I thank my committee for reading through drafts of this thesis

and for their valuable insights and comments.

My parents, who are probably the most earnest parents in educating their child,

have been so supportive during my graduate study and to them I am forever grateful. My

father would still worry about my grades. My mother worries about me staying up too

late. They both enjoy listening to every trivial detail of my every Friday Skype “report” –

what I eat, what’s on the news or simply a bad-luck day. I hope that I will soon be able to

see them more often, maybe cook meals for them as I have always boasted about – and

they are very suspicious of – my second biggest achievement in graduate school:

becoming a pretty good cook.

I thank my boyfriend Xing Yi, who has always been there to share my happiness,

sometimes tears, stress and whatever I am facing. I cherish these years we have been

through together and I look forward to the new life waiting for us.

Page 8: Topics in Multivariate Time Series Analysis: Statistical

vii

I am thankful to my officemates at ISOM 307, who have made this “fish ball”

office so lovely and feel like home. Thanks to Phuong Anh Nguyen, for those early years

when we both basically “lived” in the office, and for advice and help she is always first to

offer. Thanks to Davit Khachatryan, my academic little “brother,” whose academic

discussions, companionship to conferences, and conversation on future plans are most

appreciated. And of course I have to thank all my fellow students in the Isenberg

Management Science Program – Patrick Qiang, Zugang Liu, Tina Wakolbinger, Trisha

Woolley, Deanna Kennedy, Baris Hasdemir, Ronrapee Leelawong, and Milad Ebtehaj,

without whom there wouldn’t have been so many parties, movie nights, barbeques, and

fun.

I am immensely blessed with a large circle of friends at UMass, each of whom has

been special and dear. I thank Qiao Lan for always lending a nonjudgmental ears through

good times or bad. I thank Nikhil Malvankar, Danxiang Li and Apurv Mathur for the

exciting journey of “Bug Power.” I thank Kwong Chan for countless sweaty badminton

sessions. I thank Shaohan Hu, Jing Yang, Shuang Feng, Liping Qiu and Guang Xu, Li Cai

and Jian Du, Ming Li and Lili Cheng – and I wish I could go on to mention all the names

I feel grateful to.

Honored to be an Isenberg Scholar, I have to thank Mr. & Mrs. Isenberg for their

generous support. I have to thank Professor Bisgaard and Vice Chancellor Michael

Malone for selecting me. My understanding of how academics could influence society

has been profoundly sharpened through activities in the Isenberg Integration Program.

And of course I have always enjoyed the luxury to have some extra pocket money to keep

my sleepy days caffeinated.

Page 9: Topics in Multivariate Time Series Analysis: Statistical

viii

ABSTRACT

TOPICS IN MULTIVARIATE TIME SERIES ANALYSIS:

STATISTICAL CONTROL, DIMENSION REDUCTION, VISUALIZATION AN D THEIR BUSINESS APPLICATIONS

May 2010

XUAN HUANG

B.ENG., TSINGHUA UNIVERSITY

M.S., UNIVERSITY OF MASSACHUSETTS AMHERST

Ph.D., UNIVERSITY OF MASSACHUSETTS AMHERST

Directed by: Professor Søren Bisgaard

Most business processes are, by nature, multivariate and autocorrelated. High-

dimensionality is rooted in processes where more than one variable is considered

simultaneously to provide a more comprehensive picture. Time series models are

preferable to an independently identically distributed (I.I.D.) model because they capture

the fact that many processes have a memory of their past. Examples of multivariate

autocorrelation can be found in processes in the business fields such as Operations

Management, Finance and Marketing.

The topic of statistical control is most relevant to Quality Management. While

both multivariate I.I.D. processes and univariate autocorrelated processes have received

much attention in the Statistical Process Control (SPC) literature, little work has been

done to simultaneously address high-dimensionality and autocorrelation. In this

dissertation, this gap is filled by extending the univariate special cause chart and common

Page 10: Topics in Multivariate Time Series Analysis: Statistical

ix

cause chart to multivariate situations. In addition, two-chart control schemes are

extended to nonstationary processes. Further, a class of Markov Chain models is

proposed to provide accurate Average Run Length (ARL) computation when the process

is autocorrelated.

The second part of this dissertation aims to devise a dimension reduction method

for autocorrelated processes. High-dimensionality often obscures the true underlying

components of a process. In traditional multivariate literature, Principal Components

Analysis (PCA) is the standard tool for dimension reduction. For autocorrelated processes,

however, PCA fails to take into account the autocorrelation information. Thus, it is

doubtful that PCA is the best choice. A two-step dimension reduction procedure is

devised for multivariate time series. Comparisons based on both simulated examples and

case studies show that the two-step procedure is more efficient in retrieving true

underlying factors.

Visualization of multivariate time series assists our understanding of the process.

In the last part of this dissertation a simple three-dimensional graph is proposed to assist

visualizing the results of PCA. It is intended to complement existing graphical methods

for multivariate time series data. The idea is to visualize multivariate data as a surface

that in turn can be decomposed with PCA. The developed surface plots are intended for

statistical process analysis but may also help visualize economics data and, in particular,

co-integration.

Page 11: Topics in Multivariate Time Series Analysis: Statistical

x

TABLE OF CONTENTS Page

ACKNOWLEDGEMENTS ................................................................................................ v ABSTRACT ..................................................................................................................... viii LIST OF TABLES ........................................................................................................... xiii LIST OF FIGURES ......................................................................................................... xiv LIST OF SYMBOLS ...................................................................................................... xvii CHAPTER 1. INTRODUCTION ........................................................................................................ 1

1.1 Relevance of Multivariate Time Series Models in Business Applications ........... 3

1.1.1 Case 1: the Furnace Data ........................................................................... 4 1.1.2 Case 2: the Hog Data ................................................................................. 6 1.1.3 Case 3: the Interest Data ............................................................................ 7 1.1.4 Case 4: the Fama-French Data ................................................................... 8

1.2 An Overview of This Dissertation ...................................................................... 10

2. MODEL-BASED MULTIVARIATE MONITORING CHARTS FOR

AUTOCORRELATED PROCESSES ........................................................................ 12

2.1 Statistical Process Control .................................................................................. 12

2.1.1 Common Causes and Special Causes ...................................................... 13 2.1.2 Control Charts Constituents: Charted Data.............................................. 13 2.1.3 Control Charts Constituents: Control Limits ........................................... 14 2.1.4 Average Run Length ................................................................................ 14 2.1.5 Phase I and Phase II ................................................................................. 15 2.1.6 The Family of Control Charts .................................................................. 15

2.2 Motivation: the Ceramic Furnace Example ........................................................ 15 2.3 A Discussion on “in Control in a Broader Sense” and the Chapter Overview ... 17 2.4 Literature Review of Traditional Multivariate SPC ............................................ 19 2.5 Standard Charts Applied to Autocorrelated Data ................................................ 21 2.6 Proposed Method: The Multivariate Special and Common Cause Charts.......... 26 2.7 Revisiting the Furnace Data ................................................................................ 28 2.8 Run Length Comparisons ................................................................................... 32 2.9 Conclusion .......................................................................................................... 47

Page 12: Topics in Multivariate Time Series Analysis: Statistical

xi

3. A CLASS OF MARKOV CHAIN MODELS FOR AVERAGE RUN LENGTH COMPUTATION FOR AUTOCORRELATED PROCESSES ................................... 49

3.1 Introduction ......................................................................................................... 49 3.2 The Univariate AR(1) process............................................................................. 51 3.3 Extension to the Univariate ARMA(p, q) process .............................................. 55

3.3.1 Extension to the Univariate AR(2) Process ............................................. 56 3.3.2 Extension to the Univariate ARMA(1, 1) Process ................................... 58 3.3.3 Extension to the Univariate ARMA(p, q) Process ................................... 61

3.4 Extension to the Multivariate ARMA(p, q) process ........................................... 62 3.5 Computation Efficiency ...................................................................................... 64

3.5.1 How Fine Should the Mesh be? ............................................................... 65 3.5.2 Matrix Computation ................................................................................. 69

3.6 Conclusion and Discussion ................................................................................. 70

4. COMMON FACTORS AND LINEAR RELATIONSHIPS OF MULTIVARIATE

TIME SERIES ............................................................................................................ 72

4.1 Motivation ........................................................................................................... 72 4.2 Literature Review................................................................................................ 74

4.2.1 The PCA Method ..................................................................................... 75 4.2.2 The Box-Tiao Method .............................................................................. 75

4.2.3 PCA on aΣ ............................................................................................. 78

4.2.4 The Badcock Method ............................................................................... 78 4.2.5 Dynamic Factor Model ............................................................................ 79

4.3 Reflections on the Literature ............................................................................... 80

4.3.1 The Box-Tiao Method and Canonical Correlation Analysis .................... 80 4.3.2 The Box-Tiao Method, Cointegration and Co-movement ....................... 82 4.3.3 The Box-Tiao Method and the Badcock Method ..................................... 83 4.3.4 A Simple Augmented Model .................................................................... 84 4.3.5 A Simulation Study .................................................................................. 86 4.3.6 Other Application of the Box-Tiao Method ............................................. 91 4.3.7 Spurious Box-Tiao Components .............................................................. 92

4.4 A Combined method of both the PCA and the Box-Tiao Methods ..................... 93

4.4.1 The Problem of Using the Box-Tiao method Individually ...................... 93 4.4.2 A Two-step Combined Method ................................................................ 96 4.4.3 Demonstration of the Combined Method ................................................ 97

Page 13: Topics in Multivariate Time Series Analysis: Statistical

xii

4.4.4 Simulated Study Revisit ......................................................................... 100 4.4.5 Furnace Data Revisited .......................................................................... 103

4.5 Business Applications ....................................................................................... 107

4.5.1 Statistical Process Control ..................................................................... 107 4.5.2 Applications in Asset Pricing ................................................................. 108

5. VISUALIZING PRINCIPAL COMPONENTS ANALYSIS FOR MULTIVARIATE

PROCESS DATA .......................................................................................................112

5.1 Motivation ..........................................................................................................112 5.2 Graphical Representation of Multiple Time Series ............................................114 5.3 Principal Components Decomposition ..............................................................116 5.4 Examples ............................................................................................................118

5.4.1 Example 1: The Furnace Temperature Data............................................119 5.4.2 Example 2: The Hog Data ...................................................................... 122

5.5 Further Discussion: Adding Up the Surfaces .................................................... 126 5.6 Visualization of Comovement ........................................................................... 129 5.7 Conclusion ........................................................................................................ 131

6. CONCLUSION ......................................................................................................... 133

6.1 Future Work ...................................................................................................... 134 APPENDICES A. EIGENVALUES INVARIANT TO TRANSFORMATION ..................................... 135 B. ARL CALCULATION FOR RESIDUAL BASED METHOD ................................ 136 C. PROOF OF (3.1) ....................................................................................................... 138 D. DETERMING THE END STATES IN Figure 3-3 ................................................... 139 BIBLIOGRAPHY ........................................................................................................... 140

Page 14: Topics in Multivariate Time Series Analysis: Statistical

xiii

LIST OF TABLES

Table 2-1. ARL0 With Different set of AR Coefficient Φ for Hotelling’s 2T . ............... 22

Table 2-2. Examples examined. “++ ” means both eigenvalues are positive. ................. 35

Table 2-3. ARL’s comparison for four cases. NC= Noncentrality. “Z” is the Hotelling’s 2T chart based on the observed data, “A” is the residuals based Hotelling’s 2T chart.................................................................................................................... 36

Table 2-4. ARL’s comparison for Case 5 and 6. NC= Noncentrality. “Z” stands for observed data based Hotelling’s 2T chart, “A” stands for residuals based Hotelling’s 2T chart. ................................................................................................ 41

Table 3-1: Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on AR(1) models. ........................................... 55

Table 3-2. Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on AR(2) models. ........................................... 58

Table 3-3. Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on ARMA(1,1) models. .................................. 60

Table 3-4. Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on VAR(1) models. ......................................... 64

Table 4-1. Cointegration and Co-movement. .................................................................... 83

Table 4-2. Correlation between each factor and its matching component. ..................... 102

Table 5-1. Principal Components Analysis Based on the Covariance Matrix of the Furnace-Temperature Data .....................................................................................119

Table 5-2. Principal Components Analysis based on the Correlation Matrix of the Hog Data ....................................................................................................................... 124

Page 15: Topics in Multivariate Time Series Analysis: Statistical

xiv

LIST OF FIGURES

Figure 1-1. A large ceramic furnace. ................................................................................... 4

Figure 1-2.Hourly temperature measurements at 9 locations of the ceramic furnace. ....... 5

Figure 1-3. A traditional graphical representation of the Hog data..................................... 7

Figure 1-4. Monthly deposit interest rates in Taiwan: Mar 1961 to Jul 1989. .................... 8

Figure 1-5. Fama-French factors from Nov 1942 to Jul 2008. ........................................... 9

Figure 2-1. A typical control chart. ................................................................................... 13

Figure 2-2. Temperature measurements at 2 adjacent locations of a ceramic furnace, (a) Combined display plotted on a common scale. (b) Separate displays on separate scales. ...................................................................................................................... 16

Figure 2-3. A 2T chart using the Holmes and Mergen (1993) estimator for the covariance matrix. ..................................................................................................................... 23

Figure 2-4. Control charts on Phase II production: (a) Shewhart chart on 8z ; (b) Shewhart

chart on 9z ; (c) Hotelling’s 2T on both dimensions. The control limits derived

from Phase I data. ................................................................................................... 25

Figure 2-5. The autocorrelation of the residuals after fitting the IMA model with 5% significance limits for the autocorrelations: (a) residuals from 8z ; (b) residuals

from 9z .................................................................................................................... 29

Figure 2-6. Hotelling’s 2T on residuals after fitting the Vector IMA(1, 1) model. .......... 30

Figure 2-7. Special cause chart with 8z and fitted values 8z imposed. ............................ 31

Figure 2-8. A Time series plot (a) and a box plot (b) of tz8∇ . .......................................... 32

Figure 2-9. Triangular Region of 1φ and 2φ . .................................................................... 34

Figure 2-10. ARL’s Comparison by graphs. ..................................................................... 45

Figure 2-11. 21φ , 22φ Space when eigenvalues are 0.2 and 0.8 respectively. ................... 47

Figure 3-1. Markov Chain State Space for AR(1) model. ................................................ 52

Figure 3-2. Markov Chain State Space for AR(2) model. ................................................ 57

Page 16: Topics in Multivariate Time Series Analysis: Statistical

xv

Figure 3-3. Markov Chain State Space for ARMA(1,1) model. ....................................... 59

Figure 3-4. Markov Chain State Space of VAR(1) model. ............................................... 63

Figure 3-5. ARL v.s. s for AR(1) model, 0φ = . ............................................................... 66

Figure 3-6. ARL v.s. s for AR(1) model, 0.5φ = . ............................................................ 66

Figure 3-7. ARL v.s. s for AR(2) model, 1 20.5, 0.2φ φ= = . .............................................. 67

Figure 3-8. ARL v.s. s for ARMA(1,1) model, 0.8, 0.5φ θ= = . ..................................... 68

Figure 3-9. Sparse transition matrices. ............................................................................. 69

Figure 4-1. A traditional time series plot of temperatures measured at nine different locations of a large industrial furnace. .................................................................... 73

Figure 4-2. Time Series plot of 1y , 2y and 1n . ................................................................ 87

Figure 4-3. (a) The Time Series plot of 1y , 2y , 1n and the first two PCs; (b) The Matrix

plot of 1y , 2y and 1n against two PCs. .................................................................. 89

Figure 4-4. (a) The Time Series plot of 1y , 2y , 1n and the first three Box-Tiao

components; (b) The Matrix plot of 1y , 2y and 1n against the first three Box-Tiao

components. ............................................................................................................ 90

Figure 4-5. Predictability of Box-Tiao Factors of the Furnace data. ................................ 94

Figure 4-6. Correlation of each Box-Tiao factor with each dimension of observed data. 95

Figure 4-7. (a) The Time Series plot of 1y , 2y , 1n and the two PCA-BT components; (b)

The Matrix plot of 1y , 2y and 1n against the two PCA-BT components. ........... 101

Figure 4-8. The Four PCA-BT Factors on Furnace data: Time Series plots and Autocorrelation Fuction. ....................................................................................... 104

Figure 4-9. Correlation between the four BT-PCA factors and each dimension of the observed data. ....................................................................................................... 105

Figure 4-10. The Four PCAs on Furnace data: Time Series plots and Autocorrelation Fuction. ................................................................................................................. 106

Figure 4-11. How well do results from statistical factor retrieving methods match with theoretical proposed pricing factors. ...................................................................... 111

Figure 5-1. Surface plot of the industrial furnace temperature profile. ...........................115

Page 17: Topics in Multivariate Time Series Analysis: Statistical

xvi

Figure 5-2. Contour plot of the furnace temperature profile. ..........................................116

Figure 5-3. Plot of the Eigenvector Elements Versus Their Dimension Index Based on the Covariance Matrix of the Furnace Data with (a) for 1e , (b) for 2e and so forth. . 120

Figure 5-4. Principal component surfaces for the Furnace Temperature Data: (a) surface of the first principal component 1C , (b) surface of the second principal component

2C , (c) surface of the third principal component 3C , (d)surface of the ninth

principal component 9C . ...................................................................................... 121

Figure 5-5. A three-dimensional scatter plot of the furnace data. ................................... 122

Figure 5-6. Surface plot of the Hog Data. ....................................................................... 123

Figure 5-7. Plot of the eigenvector elements versus their dimension index based on the correlation matrix of the Hog data with (a) for 1e , (b) for 2e and so forth. ......... 124

Figure 5-8. Principal component surfaces for the Hog Data: (a) surface of the first principal component 1C , (b) surface of the second principal component 2C , (c)

surface of the third principal component 3C , (d)surface of the fourth principal

component 4C . ..................................................................................................... 125

Figure 5-9. Adding up the rank one principal component surfaces for the Hog data: (a) surface of the added-up first two principal components 2X , (b) surface of the

added-up first three principal components 3X . .................................................... 127

Figure 5-10. Adding up the rank one principal component surfaces for the Furnace data: (a) surface of the added-up first two principal components 2X , (b) surface of the

added-up first four principal components 4X , (c) surface of the added-up first six

principal components 6X , (b) surface of the added-up all principal components X .

............................................................................................................................... 128

Figure 5-11. Time series plots of the five principal components, PC1, …,PC5 for the Hog data. ....................................................................................................................... 131

Page 18: Topics in Multivariate Time Series Analysis: Statistical

xvii

LIST OF SYMBOLS

tz The multivariate time series we observe.

ta A vector of white noise, mean of which is 0 .

µ The mean of tz .

zΣ The variance-covariance matrix of tz .

aΣ The variance-covariance matrix of ta .

( )lΓ ( ),t t lCov +z z : the cross-covariance matrix of tz at lag l.

B The backshift operator.

t Time point t.

k The dimension of tz and ta .

r The dimension of common factors (true underlying factors).

p The autoregressive order.

q The moving average order.

l Lag.

Page 19: Topics in Multivariate Time Series Analysis: Statistical

1

CHAPTER 1

INTRODUCTION

Multivariate Time Series models describe relationships among a vector of k time

series variables 1tz , 2tz , …, ktz . Multivariate processes arise when several related time

series are observed simultaneously over time instead of observing just a single series as is

the case in univariate time series analysis. In the study of multivariate processes, a

framework is needed for describing not only the properties of the individual series but

also the possible cross-relationships among the series. The purposes of analyzing and

modeling the series jointly are to understand the dynamic relationships over time among

the series and to improve the accuracy of forecasts for individual series by utilizing the

additional information available from the related series in the forecasts for each series.

Among all the models, we base our discussion on the vector autoregressive integrated

moving average (VARIMA) time series models.

For a comprehensive but elementary introduction to VARIMA, see Brockwell and

Davis (2002). See also Tiao and Box (1981). For a more in-depth discussion, see Reinsel

(1997). Here we only state basic facts and define terms.

Let }{ tz be a k-dimensional vector autoregressive moving average ARMA(p,q)

time series tt BB aΘzΦ )(~)( = where µzz −= tt~ is the mean adjusted observation vector

(in the non-stationary case I define t t=z z% because the non-stationary process does not

have a mean; stationarity is to be defined below), 1( ) ppB B B= − − −Φ I Φ ΦK ,

( ) 1q

qB B B= − − −Θ I Θ ΘK are matrix-valued polynomials, I is the kk× identity matrix,

B the backshift operator and it is assumed that ~ ( )iid

t k aNa 0,Σ , see e.g. Reinsel (1997).

Page 20: Topics in Multivariate Time Series Analysis: Statistical

2

The ARMA(p, q) process is said to be strictly stationary if the probability

distribution of 1 2, , ,

nt t t z z zK and

1 2, , ,

nt l t l t l+ + + z z zK are the same for arbitrary times 1t ,

2t , …, nt , all n, and all lags and leads 1, 2,l = ± ± K It is said to be weakly stationary if all

the roots of 0)}(det{ =BΦ are greater than one in absolute value and invertible if all the

roots of 0)}(det{ =BΘ are greater than one in absolute value. In most of what follows I

call a process stationary if it satisfies the weak stationarity criteria. If not specified

separately, I also assume that the processes are invertible. Hence, provided the inverse

exists, ttBB azΦΘ =− ~)()]([ 1 . In other words, )()]([ 1 BB ΦΘ − is a linear filter that when

applied to the data tz~ produces a white noise series ta .

When the process is stationary, it possesses constant mean and variance. In this

dissertation, I use µ to denote the mean of tz , use zΣ to denote the contemporaneous

variance-covariance matrix of tz . Further I use ( )lΓ to denote the cross-covariance

matrix at lag l, between tz and t l+z . Note

( ) ( ) ( ) ( )( ) ( ), , ,t t l t l t t l t l ll Cov Cov Cov l+ + + + −

′′ ′ Γ = = = = Γ − z z z z z z .

In modeling multivariate time series processes, we generally want to describe,

predict, and control the processes. A simple example where a multivariate time series

model is useful is the daily inventory level of ice cream and waffle cone in a local

grocery store. To maintain a healthy inventory level, demand data of past years could be

examined and a time series model is fitted. It will not be surprising that the demand data

would show patterns and seasonality with peaks during hot seasons and lows during cold

seasons. Also the demands of these two complementary products would resemble each

Page 21: Topics in Multivariate Time Series Analysis: Statistical

3

other and are fairly cross-correlated. A two-dimension Seasonal Vector ARIMA model

would be capable of capturing these patterns and provides a fairly accurate short term

prediction of the demands. Actions could be taken in advance upon a predicted extreme

demand and it helps prevent supply falling short or a glut of supply eventually incurring

high inventory cost. This is only a simple example of how multivariate time series

modeling can be used to describe, predict and control a business process. More real cases

are provided in the next section.

1.1 Relevance of Multivariate Time Series Models in Business Applications

Most, if not all business applications are by nature multivariate and autocorrelated.

Multivariability is rooted in processes where more than one variable is considered

simultaneously to provide a more comprehensive picture of the processes. Time series

models are preferable to an I.I.D. (independently identically distributed) model because

they capture the fact that many processes have a memory of their past, i.e., the current

state of the processes depend on their past states to some extent. In this section, four

business cases are introduced to provide insights into why multivariate time series models

are relevant in business applications. These cases include operations management process,

econometric data and financial data. However it should be acknowledged that

multivariate time series models also can be applied in many more fields, e.g., marketing,

accounting. The cases provided here will be referred to from time to time throughout the

dissertation.

Page 22: Topics in Multivariate Time Series Analysis: Statistical

4

1.1.1 Case 1: the Furnace Data

Figure 1-1 shows a large ceramic furnace. The raw materials are poured in at the

left hand side. The materials then get heated; reactions take place along the furnace and

finally the finished materials flow out at the opposite end. One important responsibility of

operations management is to supervise the quality. Maintaining high yield is critically

important to the economy of the firm. Modern quality control has shifted from

dependence on final inspection to closely monitoring the process. For this reason, thermo

couples are installed in both the top and bottom of the furnace at equal distance along the

length of the furnace and the temperatures are sampled hourly and monitored closely.

Clearly this process is multivariate.

Raw

Materials

Finished

Materials

)1(tZ

)2(tZ

Raw

Materials

Finished

Materials

)1(tZ

)2(tZ

Figure 1-1. A large ceramic furnace.

Page 23: Topics in Multivariate Time Series Analysis: Statistical

5

Time [Hours]

1105.0

1102.5

1100.0

9060301

1173

1170

1167

1205

1204

1203

1207.2

1206.4

1205.6

1180

1179

1178

1116.5

1116.0

1115.5

9060301

1047.6

1047.0

1046.4

1040.0

1039.5

1039.0

9060301

1019.2

1018.4

1017.6

z1 z2 z3

z4 z5 z6

z7 z8 z9

Figure 1-2.Hourly temperature measurements at 9 locations of the ceramic furnace.

This process is autocorrelated as well. By eyeballing each individual series in

Figure 1-2 we find none of them resemble an I.I.D process. They all exhibit certain

dynamic and even some non-stationarity. More interestingly, 6tz through 9tz all look

alike; they reach peaks and troughs around the same time.

In this dissertation, I answer the following questions: how do we implement

control charts on multivariate time series processes? How do we judge a process is in-

control if the process is non-stationary? What is the latent factor that make 6tz through

9tz all look alike? Can we reduce the dimension of the time series to a number smaller

than nine?

Page 24: Topics in Multivariate Time Series Analysis: Statistical

6

Case 1 is a major motivation of this dissertation. Every chapter of this dissertation

aims at one perspective of understanding and controlling such a process. At the

conclusion of this dissertation, building blocks of results from each chapter will be put

together to show how the goal is achieved.

1.1.2 Case 2: the Hog Data

The second example is an econometrics example. Figure 1-3 shows the traditional

graphical representation of the so-called “hog data” discussed extensively in the

statistical time series literature first used by Quenouille (1957) and later Box and Tiao

(1977) and many others. It is a five-dimensional multiple time series consisting of 82

yearly observations from 1867-1948 of quantities relevant to the US hog trade. The five

dimensions are Hog supply, Hog price, Corn price, Corn supply and Farm wages

respectively. The data was originally provided in “Historical Statistics of the United

States 1789-1945” and “Statistical Abstract of the United States, 1950” published by the

US Department of Commerce. Quenouille (1957) and Box and Tiao (1977) made a few

adjustments where necessary, log transformed each series, lagged some and linearly

coded the logs. I use the data as adjusted by Box and Tiao (1977).

Page 25: Topics in Multivariate Time Series Analysis: Statistical

7

Time [Year]

Dat

a

19471939193119231915190718991891188318751867

1750

1500

1250

1000

750

500

Variable

x3x4x5

x1x2

Figure 1-3. A traditional graphical representation of the Hog data.

This data is again multivariate and autocorrelated. One question of interest to the

economists is whether there is any equilibrium or long term relationship among the five

series, that is, whether co-integration exists among the five series. Further more, how

does co-integration imply about underlying factors? I answer these questions in Chapter 4

and extend the idea of co-integration to co-movement. Another problem I consider is

whether there is a better way to display multivariate time series than the traditional graph

of Figure 1-3, which superimposes all the series together and is hard to read. In Chapter 5

I propose a solution – a 3D surface plot complementary to the traditional displays. I

demonstrate how this plot can assist our perception of PCA and co-integration.

1.1.3 Case 3: the Interest Data

Figure 1-4 is the time series plot of the one-, three-, and six-month deposit interest

rates in Taiwan from March 1961 to July 1989. The observations represent the averages

of the deposit rates offered by nine major banks. The two big swings are associated with

the two recessions caused partially by the oil crises around 1973 and 1981. Not

Page 26: Topics in Multivariate Time Series Analysis: Statistical

8

surprisingly, the three interest rates tend to move together. Although this is more or less

“man-made” data because the interest rate is controlled by the government, it is still

interesting to ask: are the three interest rates co-integrated? How many true common

factors are there?

Interest Rate

Year

Month

19891985198119771973196919651961

MarMarMarMarMarMarMarMar

14

12

10

8

6

4

6-month

3-month

1-month

Figure 1-4. Monthly deposit interest rates in Taiwan: Mar 1961 to Jul 1989.

1.1.4 Case 4: the Fama-French Data

Fama and French (1992) investigated important factors that explain the cross-

sectional variability of expected stock returns. They find the overall market factor, the

size factor and the book-to-market equity factor are significantly priced.

These factors were constructed based upon financial economics theories. The size

factor, SMB (small minus big), is the difference between the average returns on small-

stock portfolio and the average return on big-stock portfolio. The book-to-market equity

factor, HML (high minus low), is the difference between the average of the return on

Page 27: Topics in Multivariate Time Series Analysis: Statistical

9

high-BE/ME (book equity to market equity) portfolio and the average return on low-

BE/ME portfolio.

Since the paper was published, Dr. Kenneth French continuously updates the

factor data upon newly available stock returns and makes them available through his

website. Figure 1-5 shows the three Fama-French factors from Nov 1942 to Jul 2008.

Also available is the return on portfolios formed on each size decile (stocks are ranked

according to size – market capitalization, and then divided into ten groups; portfolios are

formed for each group).

20

10

0

-10

-20

2005198419631942

NovNovNovNov

20

10

0

-10

-20

Year

Month

2005198419631942

NovNovNovNov

10

0

-10

Rm-Rf SMB

HML

Figure 1-5. Fama-French factors from Nov 1942 to Jul 2008.

Stock market price/return is a commonly analyzed multivariate autocorrelated

data set. In this context, what I am interested is how these theory-based factors possibly

match empirically determined latent factors through some kind of dimension reduction

method. I look into this question in Chapter 4.

Page 28: Topics in Multivariate Time Series Analysis: Statistical

10

1.2 An Overview of This Dissertation

Chapter 2 through Chapter 5 each treat an individual problem yet fit in the same

scope of the dissertation topic. Chapter 2 deals with process monitoring of multivariate

processes and in particular with the autocorrelation and non-stationarity issues that

appears to be incompatible with traditional multivariate control charts and thus may

seriously impact their performance. We suggest modeling processes with multivariate

ARIMA time series models and propose two model-based monitoring charts. One

monitors the predicted value and provides information about the need for mean

adjustments. The other is a Hotelling’s T2 control chart applied to the residuals. The ARL

performance of the residual based Hotelling’s T2 chart is compared to the observed data

based Hotelling’s T2 chart for a group of first order vector autoregressive models. We

show that the new chart in most cases performs well.

Chapter 3 deals with Average Run Length computation problem for autocorrelated

processes. The ARL of conventional control charts is typically computed assuming

temporal independence. However, this assumption is frequently violated in practical

applications. Alternative ARL computations have often been conducted via time

consuming and not necessarily very accurate simulations. In Chapter 3, we develop a

class of Markov chain models for evaluating the run length performance of traditional

control charts for autocorrelated processes. We show extensions from the simplest

univariate AR(1) model to the general univariate ARMA(p, q) model, and the

multivariate VARMA(p, q) time series case. The results of the proposed method are

compared with simulation results and a discussion on computational accuracy and

efficiency is provided.

Page 29: Topics in Multivariate Time Series Analysis: Statistical

11

In Chapter 4 I take on the question of how to improve the current dimension

reduction methods for multivariate time series. In traditional multivariate literature, PCA

is the standard tool for dimension reduction and factor retrieving. For autocorrelated

process, however, PCA fails to take into account the time structure information. Thus it is

doubtful that PCA is the best choice. I review several existing reduction methods and

compare them with PCA. I propose to combine the PCA method and Box-Tiao method

(Box and Tiao, 1977) and therefore get a two-step procedure which is demonstrated to be

more efficient in retrieving underlying factors for autocorrelated multivariate processes. I

also investigate its application in econometrics and financial asset pricing context.

In Chapter 5 we suggest a simple method for visualizing the results of PCA

intended to complement existing graphical methods for multivariate time series data

applicable for process analysis and control. The idea is to visualize multivariate data as a

surface that in turn can be decomposed with PCA. The surface plots developed in this

research are intended for statistical process analysis but may also help visualize economic

data and in particular co-integration.

Chapter 2 is based on Huang and Bisgaard (2010); Chapter 3 is based on Huang

and Bisgaard (2009); Chapter 4 is a working paper; Chapter 5 is based on Bisgaard and

Huang (2008). This dissertation is largely problem motivated and emphasis is on how the

proposed methodologies can solve the problems.

Page 30: Topics in Multivariate Time Series Analysis: Statistical

12

CHAPTER 2

MODEL-BASED MULTIVARIATE MONITORING CHARTS FOR AUTOCORRELATED PROCESSES

For more than half a century, control charts were designed for processes that were

temporally independent. It was only until recent decades (Alwan and Roberts, 1988) that

autocorrelation started to be recognized to greatly impact the performance of control

charts. Complementary to existing literature, this chapter addresses the issue of

autocorrelation for multivariate processes.

Autocorrelation or nonstationarity may seriously impact the performance of

conventional Hotelling’s T2 charts. We suggest modeling processes with multivariate

ARIMA time series models and propose two model-based monitoring charts. One

monitors the predicted value and provides information about the need for mean

adjustments. The other is a Hotelling’s T2 control chart applied to the residuals. The ARL

performance of the residual based Hotelling’s T2 chart is compared to the observed data

based Hotelling’s T2 chart for a group of first order vector autoregressive models. We

show that the new chart in most cases performs well.

2.1 Statistical Process Control

It was a milestone for quality management when Shewhart (1931) introduced

control charts in the early twentieth century. This technique has been proven a simple yet

effective means of understanding data from real-world processes. The profound idea of

Shewhart’s control charts reside in two key concepts, common causes and special causes.

Page 31: Topics in Multivariate Time Series Analysis: Statistical

13

2.1.1 Common Causes and Special Causes

Common causes refer to routine variation which was called “controlled variation”

by Shewhart. A process under control only displays this consistent pattern of variation.

Shewhart attributed such variation to “chance” causes. Special causes refer to exceptional

variation (uncontrolled variation) which is characterized by a pattern of variation that

changes over time in an unpredicted manner. Shewhart attributed these unpredictable

changes in the pattern of variation to “assignable” causes. See Alwan and Roberts (1988)

for more disscussion.

2.1.2 Control Charts Constituents: Charted Data

Figure 2-1. A typical control chart.

A typical control chart is shown in Figure 2-1. Resembling a time series chart, it

charts the observations directly or a summary statistics if sub-grouping is used, e.g.,

subgroup mean, subgroup range. Sub-grouping refers to grouping the process outcomes

into small samples. It could happen naturally, e.g., taking products from the same batch

as a subgroup; or it could be done purposely because sub-grouping and taking average

make “chance” variation smaller and thus a same-size mean shift is more easily

Page 32: Topics in Multivariate Time Series Analysis: Statistical

14

detectable. In multivariate cases, charted data could also be a vector based statistics, like

the Hotelling’s T2.

2.1.3 Control Charts Constituents: Control Limits

No matter what is being charted, a set of control limits will be imposed, as the

Upper Control Limit (UCL) and Lower Control Limit (LCL) shown in Figure 2-1. The

limits function to help judge whether the process is in control. In some sense, the charting

process works like a series of hypothesis tests. When a new observation or a new charting

statistics becomes available, it will be compared against the limits. Any statistic outside

the limits is considered a sign of out-of-control and process check would be called. Tight

control limits are preferable so they are sensitive to anything abnormal; however, too

tight control limits have a high probability of making false alarms. Production frequently

interrupted by false alarms will result in unnecessary operating cost. A tradeoff has to be

made by choosing a set of effective and economical limits.

2.1.4 Average Run Length

When developing a control chart algorithm, there are two important questions

concerning the performance. First, how often will there be false alarms while nothing

actually has changed? Second, how quickly will we detect systematic change?

The answers to these two questions are similar to “Type I Error” and “power”

respectively in hypothesis testing context. In control charts context, Average Run Length

is the measure designated to answer them. By definition, ARL is how long on the average

we will plot successive control charts points before we detect a point beyond the control

Page 33: Topics in Multivariate Time Series Analysis: Statistical

15

limits. Often, the answer to the first question is called “in-control ARL,” or denoted

simply by ARL0 while the answer to the second question is called “out-of-control ARL.”

2.1.5 Phase I and Phase II

There are two phases for control charts application. The retrospective Phase I

study is a careful scrutiny of a sequence of past data. If based on a period when the

process is deemed to be “well behaved,” Phase I data provides a reference for how an

“in-control” process behaves. Statistical characteristics of the process are estimated in

this phase and are later used to determine UCL and LCL for Phase II. Normally Phase I

does not last long in real production. After the process is studied and understanding in

gained, we switch to Phase II where the purpose is to monitor the process to maintain

control.

2.1.6 The Family of Control Charts

Basic control charts include Individual Moving Range (IMR) chart, x R− chart,

x s− chart, p chart, np chart, c chart and u chart. More recently there are Cusum and

EWMA charts. For a more detailed discussion, see Montgomery (2008). This chapter is

dedicated to a multivariate chart, Hotelling’s T2 chart; more details are provided in

section 2.4.

2.2 Motivation: the Ceramic Furnace Example

With the proliferation of computers, automatic sensor technology,

communications networks and sophisticated software, data used for statistical process

control of industrial processes are increasingly multivariate (vector-valued) and sampled

Page 34: Topics in Multivariate Time Series Analysis: Statistical

16

individually at high sampling rates. Typically data vectors arrive one by one at equal

time intervals, e.g., every hour or every minute. Sub-grouping, as for example done with

the traditional Shewhart X R− control chart, is often not convenient, useful or desirable.

For such applications, the data will typically not only be cross correlated, but also

autocorrelated especially if sampled quickly relative to the dynamics of the system being

monitored. Indeed, the data stream used as input to process monitoring algorithms are

often high dimensional vector time series.

We consider Case 1 from Chapter 1 as a motivational example. The original

dimension of this problem was nine readings from nine locations, but for now, we

consider only two adjacent time series, 8z and 9z . A sample of 100 hourly temperature

observations from these two different locations of the furnace is shown in Figure 2-2.

Time [Hour]

Tem

pera

ture

1009080706050403020101

1038

1032

1026

1020

8z

9z

Time [Hour]

Tem

pera

ture

1009080706050403020101

1038

1032

1026

1020

8z

9z

Time [Hour]

100501

1040.0

1039.5

1039.0

1038.5

100501

1019.0

1018.5

1018.0

1017.5

z8 z98z 9z

Time [Hour]100501

1040.0

1039.5

1039.0

1038.5

100501

1019.0

1018.5

1018.0

1017.5

z8 z98z 9z

(a) (b)

Figure 2-2. Temperature measurements at 2 adjacent locations of a ceramic furnace, (a) Combined display plotted on a common scale. (b) Separate displays on separate scales.

From a practical point of view, the most noticeable observation from Figure 2-2(a)

is that the temperature variation is extremely small given the high level of the

temperatures and the large scale of the furnace. We further see from Figure 2-2(b) that

although the process seems to maintain a relatively stable level and show limited

variability, it exhibits patterns and signs of non-randomness. Indeed, a furnace system as

Page 35: Topics in Multivariate Time Series Analysis: Statistical

17

large as this, has natural inertia and will exhibit high autocorrelation; so do most chemical

plants. If we used the usual control charts and imposed Western Electric Rules we would

find that this process appears to lack statistical control. Yet this furnace and in particular

during this time period had very well behaved judged by the team of control engineers

who monitored it. That raises the question: is this process in control? Can non-stationary

process possibly be in control as well? Or more fundamentally, what do we mean by

statistical control? Further, if this process could be considered in control, as was insisted

by the control engineers, are the traditional control charts immediately applicable? Or

could we device a control chart more suitable?

2.3 A Discussion on “in Control in a Broader Sense” and the Chapter Overview

In the past it has often been assumed or implied that processes in a state of

statistical control should be modeled as independently and identically distributed (I.I.D.)

random variables. More recently this view has been challenged, for example, by Alwan

and Roberts (1988) and Box and Paniagua-Quiñones (2007). By quoting Shewhart who

pointed out that a state of statistical control implies predictable processes, Alwan and

Roberts (1988) argued for a more flexible view whereby a process to which a time series

model can successfully be fitted, implying predictability, should be considered “in control

in a broader sense.”

Traditional control charts are outdated for autocorrelated processes and new

charts are therefore sought after. For univariate cases, three approaches have been

proposed. This first approach focuses on adjusting parameters and control limits to

accommodate the autocorrelation. See, e.g., VanBrackle and Reynolds (1997). The

second approach introduces a moving window of observations from the univariate

Page 36: Topics in Multivariate Time Series Analysis: Statistical

18

process and treats it as a multivariate problem using multivariate control charts. See, e.g.,

Krieger, Champ and Alwan (1992), Alwan and Alwan (1994), Apley and Tsung (2002),

Jiang (2004). The third and the most popular approach, is a residual-based method, which

is, by its name, to apply traditional control charts (Shewhart chart, EWMA chart,

CUSUM chart) to residuals after fitting a time series model. This approach has been

discussed by Alwan and Roberts (1988), Harris and Ross (1991), Montgomery and

Mastrangelo (1991), Lin and Adams (1996), Vander Weil (1996), Box and Luceño (1997),

Apley and Shi (1999), Rosolowski and Schmid (2003), and Box and Paniagua-Quiñones

(2007). Evaluation of the residual-based method has been studied by Wardell, Moskowitz

and Plante (1992, 1994), Superville and Adams (1994), Runger, Willemain and Prabhu

(1995), Kramer and Schmid (1997), Lu and Reynolds (1999a, 1999b, 2001). There were

a number of other methods which do not fall into any of the abovementioned categories,

such as a modified observation-based CUSUM scheme by Atienza and Ang (2002), a

CUSUM-triggerd CUSCORE chart that utilized the information contained in the

dynamics of the fault signature as proposed by Shu et al. (2002), a multivariate control

scheme on univariate feedback-controlled process by Apley and Tsung (2002), and a

regression model-based control chart by Loredo et al. (2002).

Despite the large research effort spawned by the problem of autocorrelation, much

of the previous work was based on stationary process and very little literature has

explicitly treated nonstationarity. Moreover, although an extension of the univariate

residual-based control chart to multivariate residual-based control chart has been

proposed (see, e.g., Pan 2005), it remains unclear how practical this method is and how

well it performs. This part of my dissertation thus attempts to address these two

Page 37: Topics in Multivariate Time Series Analysis: Statistical

19

remaining questions: first, how to extend the residual-based method to time-structured

multivariate process, including the nonstationary cases; second, how the proposed control

chart performs. Further, we demonstrate the practical value of our method with a case

study.

The rest of this chapter is organized as following. In Section 2.4 we briefly review

the relevant multivariate quality control literature, and in Section 2.5 we discuss problems

with the standard Hotelling’s T2 chart applied to autocorrelated process. In Section 2.6 we

draw a parallel to the univariate counterpart of the special cause chart and propose a

residual based Hotelling’s T2 control chart for multivariate process; as a complementary

chart, we also introduce a common cause chart which provides information for process

adjustment. A revisit of the case of Furnace Data follows in Section 2.7. To further

validate the generalized special cause chart, we evaluate in Section 2.8 the proposed

special cause chart using ARL comparisons based on a comprehensive discussion of a

first order vector autoregressive time series model with varying parameters. Lastly we

provide conclusions and some final thoughts in Section 2.9.

2.4 Literature Review of Traditional Multivariate SPC

Hotelling’s T2 chart is a popular tool for monitoring multidimensional cross-

correlated processes for continued process stability originally introduced by Hotelling,

(1947); see also Hald (1951), Alt (1984) , Jackson (1985), Johnson and Wichern (2002),

Mason and Young (2002), Kourti and MacGregor (1996); see Montgomery (2008) for a

recent summary of previous work. Herein I limit my discussion to the situation where the

2T chart is applied to individual observation, i.e., subgroup sample size equals 1. In

Page 38: Topics in Multivariate Time Series Analysis: Statistical

20

general, suppose 1( , )t t tkz z ′=z K is a k-dimensional data vector. Suppose further that

),(~ Σµz k

iid

t N . Hotelling T2 at time t, is given by 2 ( ) ( )t t tT −′= − −1z z S z z where z is an

estimate of the mean and S is an estimate of the covariance matrix. In retrospective

Phase I applications nt ,,1Κ= and in prospective Phase II applications, nt > ; see

Woodall (2000) for a discussion of phases in control chart applications. Typically the

estimates z and S are based on data from Phase I. In Phase I, tz is neither independent of

the sample covariance matrix S nor the sample mean z . Tracy et al. (1992) suggests

( )22 / 1tnT n− is Beta distributed and showed that the upper control limit (UCL) for the

2T statistics should be computed as ( ) ( )2 1

1 , /2, 1 /21

k n kn n B

α−

− − −− . For Phase II application,

Tracy et al. (1992) suggests ( ) ( ) ( )2 / 1 1tn n k T k n n− + − is F distributed and showed that

the UCL for the 2T statistics is ( )( ) ( ) 111 , ,1 1 k n kk n n n n k F α

−−− −+ − − .

The estimation of the covariance matrix for individuals is an issue. In the past, it

was standard to use the unbiased Maximum Likelihood estimator (MLE) S. However,

Sullivan and Woodall (1996) discussed this and compared several other estimators,

including one suggested by Holmes and Mergen (1993) based on successive differences,

which provides a more robust estimate against a sudden mean shift in the process. We

discuss this issue below.

Besides the Hotelling’s T2 control chart, cumulative sum (CUSUM) control chart

and exponentially weighted moving average (EWMA) control chart both have

multivariate versions. Woodall and Ncube (1985), Crosier (1988) and Pignatiello and

Runger (1990) proposed various multivariate CUSUM control charts. Lowry et al. (1992)

Page 39: Topics in Multivariate Time Series Analysis: Statistical

21

presented a multivariate EWMA control chart following the same scheme as the

univariate EWMA. Wierda (1994) provides a useful overview of various methods. See

also Sullivan and Woodall (1996) for an overview and comparison of several methods.

Montgomery (2008) provides an up-to-date overview of the state of the art of

multivariate control chart methods.

It is acknowledged that all these multivariate control methods are intended for

data that comply with an assumption of temporal independence. However, if the data is

autocorrelated, most of them either do not work well, or provide misleading results.

2.5 Standard Charts Applied to Autocorrelated Data

Blindly applying traditional control charts to autocorrelated processes can be

harmful. The impact on the performance of classical SPC procedures on univariate

processes has been studies by Johnson and Bagshaw (1974), Bagshaw and Johnson

(1975), and Harris and Ross (1991). In this section we discuss briefly the impact of

autocorrelation on Hotelling’s T2 control chart.

As discussed in Section 2.1, the ARL is typically used for evaluating the

performance of a control chart method. Under the assumption of temporal independence,

the 2tT at Phase II is independent. Thus the run length has a geometric distribution with

the mean equal to 1/p, where p is the probability that an individual 2tT falls outside of the

control limits. However, if the data tz are autocorrelated, the individual’s 2tT will also be

autocorrelated. Thus the logical argument that leads to the geometric distribution is no

longer valid.

Page 40: Topics in Multivariate Time Series Analysis: Statistical

22

To illustrate what may happen, consider a two dimensional first order vector

autoregressive time series, VAR(1). Specifically, let Φ be a 22× coefficient matrix and

ttt aΦzz += −1 where it is assumed that ),(~ I0a Niid

t . Note that 0Φ = is equivalent to no

autocorrelation. To simplify we assume that zΣ is known even though only its estimate

will be so in a Phase II study; this difference will not change our overall conclusion

below. In Table 2-1 I compute the ARL0 (ARL when the process is in control) of

Hotelling’s 2T on observed data for different values of Φ corresponding to different

scenarios of autocorrelation. The UCL is fixed across all cases, chosen to give an ARL0

of 300 when the process is i.i.d.. The Markov chain method introduced in Chapter 3 is

used for ARL computation.

Table 2-1. ARL0 With Different set of AR Coefficient Φ for Hotelling’s 2T .

Φ 0 0

0 0

0 0

0 0.3

0.5 0

0 0.3

0.7 0

0 0.8

0.7 0

0 0.5

− −

0.9 0

0 0.9

− −

ARL0 300.18 367.51 338.51 406.08 422.37 534.51

Note that the Φ matrices shown in Table 2-1 are all diagonal. However, this is

not essential for our conclusion. Thus Table 2-1 indicates that it is difficult to control the

ARL0 for the Hotelling’s 2T chart without knowing the underlying model.

The ARL is not the only problem; many useful techniques developed based on

assuming temporal independence produce unreliable result for autocorrelated processes.

In particular practitioners are cautioned against blindly applying standard statistical

software; they may use parameter estimators that are seriously non-robust to temporal

dependence. For example, Figure 2-3 shows a standard Hotelling’s T2 applied to the

furnace data using the Holmes and Mergen (1993) covariance estimator, an estimator

Page 41: Topics in Multivariate Time Series Analysis: Statistical

23

robust to sudden step shift. Figure 2-3 shows numerous out-of-control signals. The UCL

is computed through Phase I formula ( ) ( )2 1

1 , /2, 1 /21 k n kUCL n n B α−

− − −= − , with 100n = ,

2k = and 1/ 500α = so that ARL0 is 500.

Time [Hour]100806040201

120

100

80

60

40

20

011.911.9

Hot

ellin

g’s

T2o

n z 8

and

z 9

Time [Hour]100806040201

120

100

80

60

40

20

011.911.9

Hot

ellin

g’s

T2o

n z 8

and

z 9

Figure 2-3. A 2T chart using the Holmes and Mergen (1993) estimator for the covariance matrix.

The Holmes-Mergen’s estimator of the variance-covariance matrix is,

( )

1

1

1

2 1

n

i iin

=

′=− ∑dS d d

, (2.1)

where 1 , 1,..., 1t t t t n+= − = −d z z are successive differences, and n is the sample

size. As indicated dS was intended to provide a robust estimate of zΣ against sudden

mean step changes, similar to that sometimes advocated in the univariate case where the

moving range is a popular estimator of the variance; see Box and Luceño (1997), page 44

for comments. We acknowledge that this estimator never was intended to be used for

autocorrelated processes. However, in the present case of positive autocorrelation, the

most common type for continuously monitored processes, dS seriously underestimates

the variance-covariance matrix, resulting in the excessive number of false alarms.

Page 42: Topics in Multivariate Time Series Analysis: Statistical

24

A simple ad hoc solution to the problem of positive serial correlation is to inflate

the variance-covariance matrix by an appropriate scalar factor using an empirical

reference distribution for benchmarking. For example, a test data set of appropriate length

from a time segment where the process is deemed stable by knowledgeable process

engineers can be used as a reference distribution. The process engineers can then either

adjust the control limits to make sure no false alarms are recorded during this period or

inflate the variance-covariance matrix by a scalar factor to accomplish the same. Both

approaches are simple and work in practice; see also Kourti and MacGregor (1996).

However, this remains an ad hoc approach.

Another way to fix the problem of autocorrelation is to use the MLE, S directly.

That may work reasonably well for Hotelling’s 2T when we a have a sufficiently long

time series for the Phase I estimation of the covariance matrix; see Johnson and

Langeland (1991).

It is even more dangerous to use Hotelling’s 2T when the time series data is non-

stationary. Box and Luceño (1997) pointed out that most real world processes are non-

stationary because of the second law of thermodynamics. The furnace data, a

representative of an industry process, provides good evidence for their argument.

Analyzing the Phase I data, we find that the furnace temperature exhibit patterns modeled

well by a Vector IMA(1,1) model. If the data from the initial 100 observations shown in

Figure 2-2 are assumed to be i.i.d. normal and used as a Phase I period, we would have

found the process hopelessly out of control in the following three month, as showed in

Figure 2-4. The UCL of Individual chart (Figure 2-4a and Figure 2-4b) is determined by

the traditional formula of ˆ3µ σ+ where µ and σ are computed from Phase I, the first

Page 43: Topics in Multivariate Time Series Analysis: Statistical

25

100 observations; the LCL is determined by ˆ ˆ3µ σ− . The UCL of Hotelling’s 2T (Figure

2-4c) is computed through Phase II formula ( )( ) ( ) 111 , ,1 1 k n kk n n n n k F α

−−− −+ − − , with

1274n = , 2k = and 1/ 500α = so that ARL0 is 500.

8z

Time [Hour]120010008006004002001

1044

1042

1040

1038

1036

1034

1032

1040.571040.57

1038.631038.63

8z

Time [Hour]120010008006004002001

1044

1042

1040

1038

1036

1034

1032

1040.571040.57

1038.631038.63

9z

Time [Hour]120010008006004002001

1022

1020

1018

1016

1014

1012

1010

1019.261019.26

1017.341017.34

9z

Time [Hour]120010008006004002001

1022

1020

1018

1016

1014

1012

1010

1019.261019.26

1017.341017.34

(a) (b)

Time [Hour]120010008006004002001

800

600

400

200

0 12.512.5Hot

ellin

g’s

T2

on

z 8a

nd z 9

Time [Hour]120010008006004002001

800

600

400

200

0 12.512.5Hot

ellin

g’s

T2

on

z 8a

nd z 9

(c)

Figure 2-4. Control charts on Phase II production: (a) Shewhart chart on 8z ; (b) Shewhart

chart on 9z ; (c) Hotelling’s 2T on both dimensions. The control limits derived from

Phase I data.

Specifically, we see from Figure 2-4 (a) and (b) that both 8z and 9z exhibit a step

decline around observation 320 and then climb back up around observation 1170. Most

likely both process shifts are the effect of process adjustments by the process engineers.

Indeed temperature changes of an inherently non-stationary process is normal and should

not necessarily be a cause for alarm. Rather it is a cause for a routine adjustment, for

example, production switch from making clear glass to making tinted glass. Alarms

Page 44: Topics in Multivariate Time Series Analysis: Statistical

26

should be reserved for something extraordinary, a systems change shown up in the

residual chart.

2.6 Proposed Method: The Multivariate Special and Common Cause Charts

Echoing earlier work by Box and Jenkins (1968), Box, Jenkins and MacGregor

(1974), and Berthouex, Hunter and Pallesen (1978), Alwan and Roberts (1988) suggested

that we consider monitoring two charts, a “common cause” chart and a “special cause”

chart. The common cause chart is a chart of the fitted values or one-step-ahead forecasts

from a time series model. The special cause control chart is a time series plot on the

residuals. The common cause chart provides an estimate of the current mean assuming

the process continues to follow the same stationary or non-stationary time series model as

in Phase I, whereas the special cause chart of the residuals triggers an alarm whenever a

deviation from our expectation is too large, indicating that a special cause or system

change may have occurred. The same philosophy has been expressed by Box (1991),

Box and Luceño (1997), and Box and Paniagua-Quiñones (2007).

Based on fitting a multivariate time series model, we extend the univariate special

cause and common cause chart to multivariate horizon.

To construct a special cause chart we suggest computing the following scalar

quantity, 2 1ˆ ˆ ˆ , 1,t t a tT t−′= =a S a K where ˆa a=S Σ obtained by fitting an appropriate time

series model to }{ tz and tt BB zΦΘa ~)(ˆ)](ˆ[ˆ 1−= is the filtered data. This residual-based 2T

chart is generalization of the univariate special-cause chart proposed by Alwan and

Roberts (1988) and Box and Paniagua-Quiñones (2007).

Our proposed procedure for monitoring a multivariate process is as follows:

Page 45: Topics in Multivariate Time Series Analysis: Statistical

27

Step 1. (Phase I): Select a set of data from a time period that by the operators

familiar with the process is deemed “well behaved.” The data size n should typically be

larger than what is usually required for temporally independent data. How large depends

on how autocorrelated the process is. For moderate positive autocorrelation a minimum

of 100 seems a good rule of thumb but larger sample size is often needed. This data set

should then be plotted in all possible ways to check for outliers and other unusual

patterns. If needed we may also apply transformations to obtain approximate multivariate

normality; see Johnson and Wichern (2002).

Step 2. (Phase I): After the preliminary screening of the data we proceed with the

iterative time series modeling scheme of identification, estimation and checking

advocated by Box and Jenkins (1976). Tiao and Box (1981) and Peña, Tiao and Tsay

(2001, Chapter 11) provide good overviews of the model building process for

multivariate time series.

Step 3. (Phase I): If a time series model fits the data well, the Phase I data can be

considered to represent a predictable process and we proceed to Step 4. If not, it would be

an indication that the process is not in statistical control according the extended definition

of being predictable. Thus we need to investigate why the process is not predictable.

Step 4. (Phase II): Apply the filter tt BB zΦΘa ~)(ˆ)](ˆ[ˆ 1−= to new incoming Phase II

data. Construct a chart for tracking ,...1,ˆ 2 += ntTt . This is the special cause chart. Points

falling above the UCL indicate the advent of special causes resulting in large deviations

from the expected value. A special cause is a fundamental change to the system

generating the time series. Hence it may either be a change in the distribution of ta , or to

Page 46: Topics in Multivariate Time Series Analysis: Statistical

28

the structure of time series model specified by )(BΦ and )(BΘ and the parameters in

those polynomials. 2χ distribution is suggested for control limit computation.

Step 5. (Phase II): In addition to the special cause chart, we suggest charting each

dimension of the fitted value 1ˆ t+z either individually or jointly. These are common cause

charts. Note that for these charts we do not superimpose control limits. Rather we may

impose action limits to indicate when the predicted vector has moved too far from the

target and the process is in need of an adjustment.

With this procedure aside, let us go back to our previous question: can a

nonstationary process possibly in control or not? This question is critically important

because many real-world processes, such as our Furnace, exhibit nonstationary patterns.

Keeping in mind Alwan and Roberts’s (1988) argument on “in control in a broader

sense,” we answer this question with a resounding Yes. As long as the nonstationary

process can be fitted by a time series model and follows the same model throughout

Phase II, the only difference of controlling this process is to have a common cause chart

that behaves nonstationary. The special cause chart on the residuals will not sound an

alarm as long as the process does not meander away from what the nonstationary model

could explain.

2.7 Revisiting the Furnace Data

We now use the furnace data introduced in Section 1.1.1 and plotted in Figure 2-2,

as our Phase I data. The primary purpose of a retrospective study of the data, is to

understand the process and test whether the process can be considered in control here

interpreted in the wider Shewhart sense of being predictable. For the current data we

Page 47: Topics in Multivariate Time Series Analysis: Statistical

29

found that a Vector IMA(1,1) model fit the data well and that the residuals are well

behaved. The fitted model is,

( )-0.123 -0.358

1-0.457 -0.097t tB I B

− = −

z a

The autocorrelation functions (ACF) of the residuals after fitting a vector

IMA(1,1) model to the first 100 observation shown in Figure 2-5. The ACF’s indicates

that this model fits well for the Phase 1 data. Thus we proceed to Phase II.

Lag

Aut

ocor

rela

tion

2420161284

1.0

0.5

0.0

-0.5

-1.0

Lag

Aut

ocor

rela

tion

2420161284

1.0

0.5

0.0

-0.5

-1.0

(a) (b)

Figure 2-5. The autocorrelation of the residuals after fitting the IMA model with 5% significance limits for the autocorrelations: (a) residuals from 8z ; (b) residuals from9z

In Phase II, residuals and fitted values need to be computed recursively as time

passes using the model and fixed parameter values as estimated in Phase I. Figure 2-6

shows the residual based special cause chart for the Phase II data. The UCL is determined

through ( )2 1 ,UCL kχ α= − , where 2k = and 1/ 500α = so that ARL0 is 500. Note the

2( )kχ approximation is very rough because none of ˆ ( )BΦ , ˆ ( )BΘ and ˆ aΣ is perfect

estimate. We suspect this is a major reason that causes some false alarms and do consider

a more accurate UCL part of our future work. For now we would recommend ignoring

slightly out-of-control points because UCL has been underestimated. Although not

Page 48: Topics in Multivariate Time Series Analysis: Statistical

30

perfect, compared with Figure 2-4(c), this chart provides an impression more consistent

with how the process engineers considered this period of time.

Time [Hour]120010008006004002001

160

120

80

40

012.412.4

Hot

ellin

g’s

T2o

n R

esi

dua

ls

Time [Hour]120010008006004002001

160

120

80

40

012.412.4

Hot

ellin

g’s

T2o

n R

esi

dua

ls

Figure 2-6. Hotelling’s 2T on residuals after fitting the Vector IMA(1, 1) model.

As already indicated, the engineers from time to time adjusted the process during

the Phase II period, mostly, we surmise, by eyeballing something akin to the fitted values.

Such adjustments are natural and sometimes necessitated because of materials changes,

environmental changes, and occasional momentary power failures. Figure 2-7 overlays

8z , the common cause charts of the fitted values 8z ( 9z would be similar and thus omitted)

and the special cause chart of Hotelling’s 2T on the residuals. Both 8z and 8z are scaled

so that they can be shown and compared to the special cause chart. On the special cause

chart we see three large spikes at time 349, 520 and 1254 indicating out-of-control

situations. Around these three time points, as indicated by the three bold windows, the

process experienced a sharp plunge, a noise spike and a sudden increase. The first and

third significant alarms most likely occurred when the engineers decided to adjust the

process to a new target temperature, evidenced by the steep trends around those points.

Such adjustments are a normal part of operating of the process. With open

Page 49: Topics in Multivariate Time Series Analysis: Statistical

31

communication between the process control engineers and the quality control engineers,

each party should know the reason and not be alarmed. Only the second alarm around

observation 520 looks like a special cause. The fitted value at this point is consistent

with neither of its neighbors. Given the large inertia of the furnace, a more than one-

degree temperature drop in one hour and an immediate return in the next hour is unusual.

Thus this is more likely an outlier possibly caused by a thermo couple misfire. In fact if

we plot 8z∇ , the first difference of 8z , as shown in Figure 2-8, the spike captured in

second window of Figure 2-7 results in a first difference significantly outside its

empirical reference distribution, as circled in both charts of Figure 2-8. This further

indicates that the second special cause might be due to an outlier.

Time [Hour]

Dat

a

120010008006004002001

500

400

300

200

100

0 12.412.4

Ho

telli

ng’

sT

2on

Res

idu

als

8z

8z

Time [Hour]

Dat

a

120010008006004002001

500

400

300

200

100

0 12.412.4

Ho

telli

ng’

sT

2on

Res

idu

als

8z

8z

Figure 2-7. Special cause chart with 8z and fitted values 8z imposed.

Page 50: Topics in Multivariate Time Series Analysis: Statistical

32

12009006003001

0.4

0.0

-0.4

-0.8

-1.2

0.5

0.0

-0.5

-1.0

(a) (b)

Figure 2-8. A Time series plot (a) and a box plot (b) of tz8∇ .

The common cause chart of 8z in Figure 2-7 serves as a compass for the control

engineers. Depending on it, they are able to know where the process is relative to a given

target temperature, when it needs to be adjusted, and by how much. The same common

cause chart should be made for 9z . However, due to the high correlation between 8z and

9z , it is not shown. When the fitted model is stationary and the process is of high

dimension, a Hotelling’s 2T like chart (a generalized Euclidean distance measures) of

fitted values is recommended. Note the variance-covariance matrix of the fitted process

needs to be used, and it is not defined for non-stationary processes.

2.8 Run Length Comparisons

For non-stationary process, the Hotelling’s 2T chart applied directly to the

observed data is neither defined nor meaningful. Thus we confine the ARL performance

comparison for the Hotelling’s 2T charts to stationary process. The performance of the

special-cause chart introduced by Alwan and Roberts (1988), the univariate counterpart to

our proposed generalization, has been discussed by Wardell et al. (1992, 1994). For

univariate process, they showed that the residual-based special cause chart is not

Page 51: Topics in Multivariate Time Series Analysis: Statistical

33

necessarily the best chart to use for every type of autocorrelation. In general, the residual

chart outperforms the Shewhart chart only when the data are negatively autocorrelated.

Unfortunately, this is less common in industrial practice, making the residual-based chart

less appealing for practical use. Should the same be true for the multivariate extension,

there may be little motivation for the use of the proposed residual-based Hotelling’s 2T .

However, as we show below, the multivariate case is more complicated. Specifically, the

performance of the residual based Hotellings 2T depends on the parameter matrices as

well as the direction and size of a mean shift.

To investigate the complexity of the situation consider a 2-dimension VAR(1)

model 1( )t t t t t−− = − +z µ Φ z µ a where )(~ aΣ0,a Niid

t . Without loss of generality, we

assume 2 2a ×=Σ I . The structure of this time series model depends entirely on Φ with zΣ

being determined through Φ (Reinsel 1997, p. 30). The situation can therefore be

analyzed by investigating the performance of the two Hotelling’s 2T charts for different

values of Φ and the combination of different size and direction of the mean shift. We

base our discussion on eigenvalues of Φ , which are invariant to transformation on tz ,

see APPENDIX A. Let

11 12

21 22

φ φφ φ

=

Φ

.

The two eigenvalues are

21 1 2

1,2

4

2

φ φ φλ

± +=

where, 1 11 22φ φ φ= + and 2 12 21 11 22φ φ φ φ φ= − .

Page 52: Topics in Multivariate Time Series Analysis: Statistical

34

For stationarity we require that 2,1,1 =< iiλ ; see Reinsel, p. 29. Thus 1φ and 2φ

must satisfy: 2 1 1φ φ+ < , 2 1 1φ φ− < and 21 1φ− < < . This region is shown in Figure 2-9.

(1)(2)

(3)

(4)

φ1

φ 2

Figure 2-9. Triangular Region of 1φ and 2φ .

The stationary region can be further divided into four sub-regions according to

whether the value of the eigenvalues 2,1, =iiλ : (1) both positive, (2) both negative, (3)

one positive one negative or (4) two imaginary eigenvalues. Note with eigenvalues fixed,

Φ is not fixed. Representative examples of these four cases are presented in Table 2-2 as

Case 1 to Case 4. For each of these four cases, we examine examples of the Phase II ARL

performance under a set of mean shift scenarios given by

t

t T

t T

<=

δ

where )sin,(cos ′⋅= ααaδ , / 6iα π= with 0,1,...5i = (due to the symmetric

density function of tz%, we need not consider cases from 6i = to 11), and the scalar a

varies so that the non-centrality parameter 1y−′δ Σ δ takes values of 0.5j with .9,,1,0 Κ=j

In other words, a and α represent the size and direction of the shift vector, respectively.

Pan (2005) made a thorough discussion on three types of parameter shift without

providing ARL computation.

Page 53: Topics in Multivariate Time Series Analysis: Statistical

35

Table 2-2. Examples examined. “++ ” means both eigenvalues are positive.

Case 1: ++ 2: −− 3: −+ 4: imaginary 5: ++ 6: ++

1,2φ 1, 0.16− 0.5, 0.031− − 0.334,0.334− 0.833, 0.667− 1, 0.16− 1, 0.16−

1,2λ 0.2,0.8 0.073, 0.427− − 0.768,0.435− 0.417 0.702i± 0.2,0.8 0.2,0.8

Φ 0.5 0.3

0.3 0.5

0.25 0.25

0.125 0.25

− −

0.167 0.5

0.5 0.5

0.5 0.667

0.75 0.333

0.5 1

0.09 0.5

0.692 0.048

1.115 0.308

Calculating the ARL for Hotelling’s 2T chart based on observed data is not

straight forward because 2tT is autocorrelated. Our computation is through the Markov

chain method discussed in Chapter 3 of this dissertation. Computing the ARL for the

residuals-based Hotelling 2T chart is discussed in APPENDIX B.

Table 2-3 provides ARL comparisons for Case 1 through Case 4. The control

limits are calibrated to make ARL0 approximately 300. For each case and same mean

shift, the ARL is presented in bold face whenever the residual-based chart performs better

than the observed data chart. It can be seen that in Case 1, the residual-based chart has

larger ARL’s for all scenarios. In Case 2, the situation is the opposite with the residual-

based chart performing significantly better. The same is true for Case 4. In Case 3, the

residual-based chart performs mostly better except for shifts in directions around 6/π .

These four cases obviously do not provide a base for general inference. However,

they suggest that the residual based chart has potentially better performance in some

realistic scenarios. Indeed this conclusion is consistent with other cases we experimented

with but not shown here.

Page 54: Topics in Multivariate Time Series Analysis: Statistical

36

Table 2-3. ARL’s comparison for four cases. NC= Noncentrality. “Z” is the Hotelling’s 2T chart based on the observed data, “A” is the residuals based Hotelling’s 2T chart.

(a) Case 1

NC 0 6

π

3

π

2

π

2

3

π

5

6

π

Z A Z A Z A Z A Z A Z A 0 300.40 300.00 300.40 300.00 300.40 300.00 300.40 300.00 300.40 300.00 300.40 300.00

0.5 113.47 163.34 133.51 226.73 106.81 226.73 113.32 163.34 133.54 144.29 106.51 144.29 1 62.16 103.97 79.29 175.04 56.49 175.04 62.36 103.97 79.05 86.48 56.67 86.48

1.5 39.68 71.80 53.06 136.41 35.69 136.41 39.87 71.80 53.00 57.76 35.62 57.76 2 27.77 52.14 38.51 106.60 24.59 106.60 27.80 52.14 38.33 41.21 24.60 41.21

2.5 20.42 39.17 29.27 83.21 18.08 83.21 20.67 39.17 29.24 30.76 18.04 30.76 3 15.77 30.15 23.12 64.74 13.88 64.74 15.82 30.15 23.09 23.73 13.81 23.73

3.5 12.70 23.64 18.56 50.16 10.96 50.16 12.57 23.64 18.58 18.79 10.99 18.79 4 10.30 18.81 15.27 38.69 8.90 38.69 10.28 18.81 15.29 15.18 8.94 15.18

4.5 8.57 15.14 12.83 29.71 7.47 29.71 8.53 15.14 12.92 12.48 7.46 12.48

Continued on next page

Page 55: Topics in Multivariate Time Series Analysis: Statistical

37

Continued

(b) Case 2

NC 0 6

π

3

π

2

π

2

3

π

5

6

π

Z A Z A Z A Z A Z A Z A 0 300.90 300.00 300.90 300.00 300.90 300.00 300.90 300.00 300.90 300.00 300.90 300.00

0.5 107.37 64.20 108.61 92.22 107.68 92.66 107.27 65.90 107.46 46.48 107.53 45.49 1 56.72 28.25 56.71 45.67 56.71 45.99 56.57 29.23 56.61 18.87 56.71 18.38

1.5 34.75 16.01 35.13 27.34 35.23 27.56 35.01 16.63 34.95 10.45 34.94 10.17 2 23.57 10.45 23.83 18.24 23.70 18.41 23.68 10.87 23.93 6.86 23.65 6.67

2.5 17.06 7.48 17.19 13.09 17.11 13.22 17.12 7.79 17.15 5.02 17.02 4.89 3 12.91 5.72 13.05 9.90 12.84 10.01 12.85 5.96 13.01 3.95 12.84 3.86

3.5 10.03 4.60 10.23 7.80 9.97 7.89 10.09 4.79 10.20 3.28 9.98 3.21 4 8.09 3.84 8.23 6.34 8.02 6.42 8.04 3.99 8.23 2.84 7.99 2.78

4.5 6.67 3.30 6.84 5.30 6.60 5.36 6.69 3.43 6.81 2.52 6.54 2.47

Continued on next page

Page 56: Topics in Multivariate Time Series Analysis: Statistical

38

Continued

(c) Case 3

NC 0 6

π

3

π

2

π

2

3

π

5

6

π

Z A Z A Z A Z A Z A Z A 0 300.82 300.00 300.82 300.00 300.82 300.00 300.82 300.00 300.82 300.00 300.82 300.00

0.5 109.73 86.93 107.83 184.11 109.55 73.33 109.65 18.17 108.40 9.98 110.09 21.15 1 59.81 41.99 58.03 125.87 59.43 33.45 59.74 6.66 58.67 3.92 58.99 7.75

1.5 37.82 24.71 36.76 91.56 37.54 19.16 37.76 3.90 36.79 2.61 37.67 4.44 2 25.87 16.26 25.36 69.33 25.83 12.46 25.85 2.85 25.09 2.13 25.80 3.16

2.5 18.76 11.52 18.30 54.00 18.72 8.81 18.75 2.34 18.34 1.89 18.86 2.54 3 14.22 8.61 13.85 42.94 14.14 6.62 14.33 2.05 13.89 1.74 14.25 2.18

3.5 11.12 6.71 10.73 34.70 11.05 5.21 11.01 1.86 10.78 1.64 11.00 1.96 4 8.76 5.40 8.60 28.38 8.73 4.24 8.83 1.73 8.54 1.55 8.78 1.80

4.5 7.17 4.47 6.95 23.46 7.14 3.56 7.18 1.63 6.97 1.47 7.11 1.68

Continued on next page

Page 57: Topics in Multivariate Time Series Analysis: Statistical

39

Continued

(d) Case 4

NC 0 6

π

3

π

2

π

2

3

π

5

6

π

Z A Z A Z A Z A Z A Z A 0 301.08 300.00 301.08 300.00 301.08 300.00 301.08 300.00 301.08 300.00 301.08 300.00

0.5 109.11 47.18 107.87 41.11 109.10 37.67 109.84 38.92 107.71 45.27 109.30 50.40 1 59.17 18.65 58.03 15.87 59.17 14.30 59.61 14.74 58.01 17.53 59.00 20.05

1.5 37.43 9.84 36.66 8.42 37.32 7.57 37.49 7.71 36.71 9.08 37.41 10.47 2 25.81 6.08 25.20 5.30 25.75 4.79 25.72 4.80 25.13 5.50 25.76 6.33

2.5 18.75 4.16 18.20 3.72 18.65 3.39 18.72 3.34 18.34 3.70 18.75 4.23 3 14.18 3.07 13.81 2.83 14.09 2.60 14.24 2.52 13.78 2.70 14.06 3.04

3.5 11.05 2.40 10.73 2.27 10.95 2.12 11.04 2.02 10.68 2.10 10.92 2.33 4 8.81 1.97 8.57 1.91 8.72 1.80 8.80 1.70 8.57 1.72 8.73 1.88

4.5 7.14 1.68 6.92 1.66 7.11 1.58 7.17 1.49 7.01 1.48 7.13 1.59

Page 58: Topics in Multivariate Time Series Analysis: Statistical

40

It is interesting to note that the residual-based method is not favorable in Case 1

when both eigenvalues are positive. This case would appear to be reminiscent of the

univariate AR(1) model. For the univariate model the eigenvalue can be interpreted as the

autocorrelation. For the univariate case Wardell et al. (1992, 1994) showed that with

positive autocorrelation the special cause chart (residual-based chart) performs poorly

relative to a traditional Shewhart chart on the observed data. However, it would be rash to

conclude that residual-based Hotelling’s 2T always will perform poorly when both

eigenvalues are positive. Indeed this inference is wrong. To show this we provide two

counter examples. In both case we fix 1 1φ = and 2 0.16φ = − . The eigenvalues are the

same as in Case 1. However we chose two different realizations of Φ , shown as Case 5

and 6 in Table 2-2.

Table 2-4 shows ARL comparisons for Case 5 and Case 6. We see that in both

cases the residual-based Hotelling’s 2T performs better for most mean shift scenarios.

For example for Case 5, in some directions such as / 6π the residual-based method has a

larger ARL when there a small shift. However, as the size of the shift increases, the ARL

decreases quickly and even becomes smaller than the observed data based method. This

situation echoes what Wardell et al. (1992, 1994) found for univariate case, namely that

the residual based method is more sensitive to large shifts.

Page 59: Topics in Multivariate Time Series Analysis: Statistical

41

Table 2-4. ARL’s comparison for Case 5 and 6. NC= Noncentrality. “Z” stands for observed data based Hotelling’s 2T chart, “A” stands for residuals based Hotelling’s 2T chart.

(a) Case 5

NC 0 6

π

3

π

2

π

2

3

π

5

6

π

Z A Z A Z A Z A Z A Z A 0 300.45 300.00 300.45 300.00 300.45 300.00 300.45 300.00 300.45 300.00 300.45 300.00

0.5 113.60 102.99 113.98 223.97 99.88 104.01 114.19 81.01 113.66 73.86 100.12 73.74 1 62.49 50.12 62.54 159.72 52.70 53.29 62.27 38.30 61.64 33.92 52.43 33.63

1.5 39.79 27.62 39.51 108.03 32.57 32.01 39.93 22.38 39.42 19.57 32.62 19.20 2 27.71 16.19 27.48 69.51 22.20 21.03 27.79 14.75 27.71 12.85 22.31 12.43

2.5 20.44 9.88 20.37 42.86 16.18 14.65 20.54 10.52 20.37 9.19 16.22 8.74 3 15.88 6.25 15.71 25.54 12.34 10.64 15.80 7.95 15.71 6.99 12.33 6.52

3.5 12.62 4.11 12.49 14.88 9.79 7.99 12.57 6.28 12.43 5.56 9.76 5.09 4 10.27 2.84 10.19 8.60 7.92 6.17 10.25 5.13 10.19 4.59 7.96 4.11

4.5 8.55 2.09 8.42 5.05 6.58 4.87 8.53 4.31 8.41 3.90 6.63 3.43

Continued on next page

Page 60: Topics in Multivariate Time Series Analysis: Statistical

42

Continued

(b) Case 6

NC 0 6

π

3

π

2

π

2

3

π

5

6

π

Z A Z A Z A Z A Z A Z A 0 300.07 300.00 300.07 300.00 300.07 300.00 300.07 300.00 300.07 300.00 300.07 300.00

0.5 103.48 69.88 150.38 97.70 98.27 235.86 103.05 70.37 149.71 59.71 97.97 62.05 1 54.89 31.48 93.27 48.35 51.20 162.30 54.59 30.52 93.30 25.67 50.98 27.07

1.5 34.26 17.98 64.89 28.14 31.67 98.77 34.46 16.29 64.54 14.38 31.72 15.32 2 23.60 11.73 47.92 17.90 21.64 54.39 23.63 9.72 48.03 9.32 21.62 10.02

2.5 17.29 8.36 37.27 12.06 15.71 27.75 17.36 6.25 37.14 6.65 15.71 7.20 3 13.18 6.34 29.41 8.47 12.01 13.49 13.26 4.26 29.70 5.07 11.98 5.54

3.5 10.49 5.04 24.19 6.15 9.50 6.49 10.49 3.07 24.13 4.07 9.52 4.48 4 8.53 4.16 19.94 4.61 7.71 3.30 8.58 2.33 19.78 3.39 7.71 3.77

4.5 7.09 3.53 16.92 3.55 6.45 1.92 7.08 1.86 16.81 2.92 6.43 3.26

Page 61: Topics in Multivariate Time Series Analysis: Statistical

43

To visualize the ARL comparisons, we have summarized part of the data from

Table 2-3 and Table 2-4 in Figure 2-10. It is shown in polar coordinates how the ARL

changes as a function of the shift size and direction. The solid circles represent the

observed data based method and the dashed circles the residual-based method. Only cases

with non-centrality parameters 0 and 0.5 are shown. Take Figure 2-10(a) on Case 1 for

example. When the non-centrality parameter is 0 (no mean shift), the ARL’s of both

methods are 300 no matter of the direction of the shift. Thus the solid and dashed circles

overlap and the radius is 300, shown as the outer circle. Point A on this circle,

( )300, / 3π , represents that ARL equals 300 when a shift takes place at angle / 3π with

non-centrality 0 (essentially no mean shift). When the non-centrality parameter increases

to 0.5, implying a mean shift, both circles shrink. When the shift direction equals / 3π ,

the observed data based method shrinks to C (on the solid circle) while the residual-based

method shrinks to B (on the dashed circle). For Case 1, the solid circle shrinks more than

the dashed circle. This means the raw data based method generally has smaller ARL’s

regardless of the shift direction. However, in Case 2 and 4, the situation is the opposite.

Case 3, 5 and 6 provide more interesting scenarios. The dashed circles do not shrink

evenly in all directions. In fact, in some directions the radial distance of the residual-

based method is much shorter than the observed data based method. In other directions,

the ARL’s for the residual-based method do not shrink as much as the observed data

based method. Another interesting feature of the observed data based method is that it

appears to be more robust to the change in the shift direction, indicated by the fact that

the solid line shrinks more evenly in different directions. For the residual-based method,

it is not so. Here changes in the direction of the shift influences the ARL much more.

Page 62: Topics in Multivariate Time Series Analysis: Statistical

44

Note also that the polar diagrams in Figure 2-10 are made by connecting together the 12

points 6/π apart. If more points were used the curves would be smoother.

A

0

B

C

π 3

0

(a) Case 1 (b) Case 2

0

0

(c) Case 3 (d) Case 4

Continued on next page

Page 63: Topics in Multivariate Time Series Analysis: Statistical

45

Continued

0

0

(e) Case 5 (f) Case 6

Figure 2-10. ARL’s Comparison by graphs.

The difference in appearance between Case 1, Case 5 and Case 6 is noteworthy.

They all have the same eigenvalues, but the performance of the residual-based method

differs significantly. By fixing 1 1φ = and 2 0.16φ = − the parameters 11φ , 12φ , 21φ , 22φ lie

on a two-dimension surface in 4ℜ . To make the problem more tractable, we assume each

of the parameters remains within [ 2,2]− . This is reasonable assumption since parameters

outside this region would be unusual in practice. And, since [ ]22 111 2,2φ φ= − ∈ − , it is

inferred that [ ]11 1,2φ ∈ − . Because ( )11 22 11 111φ φ φ φ= − is symmetric around 11 0.5φ = , we

may further confine our discussion to [ ]11 0.5,2φ ∈ . Now suppose we fix 11φ . Through

1 11 22 1φ φ φ= + = and 2 12 21 11 22 0.16φ φ φ φ φ= − = − , 22φ is also fixed and so are 11 22φ φ and

12 21φ φ . Thus for each fixed value of 11φ , the feasible region is a pair of hyperbola in 12φ -

21φ coordinates. As shown in Figure 2-11, the two pairs in quadrant I and III represents

11 0.5φ = and 11 0.7φ = respectively (moving inward), the pair overlaid on the axes

Page 64: Topics in Multivariate Time Series Analysis: Statistical

46

represents 11 0.8φ = and the pairs in quadrant II and IV represent 11 0.9φ = , 11 1.1φ = ,

11 1.3φ = , 11 1.5φ = , 11 1.7φ = , 11 1.9φ = , 11 2φ = respectively (moving outward). The two

pairs of accentuated hyperbola represent the boundaries of possible values of 12φ and 21φ .

Note they are smaller than [ ] [ ]2,2 2,2− × − because of the two restrictions. Each point in

the 12φ - 21φ coordinates is mapped into a identical Φ matrix (with 12φ and 21φ fixed,

11φ and 22φ can be determined through 1 1φ = and 2 0.16φ = − ).

To explore the effect of how these parameters affect the performance of the

residual-based method, we use solid lines to represent regions where the residual based

method is competitive; by “competitive” we mean performing at least as well as the raw

data based method in some kind of shifts. We use dashed line for the converse. It is

interesting to note that for fixed 11φ and 22φ (in one pair of hyperbola), if either 12φ or 21φ

is large (larger than 0.6~0.7 generally), the residual-based method performs better. If

either 11φ and 22φ are larger than 0.9, no matter what the values of 12φ or 21φ are, the

residual based method performs better. This explains the difference between Case 1, Case

5 and Case 6.

Page 65: Topics in Multivariate Time Series Analysis: Statistical

47

−2 −1 0 1 2

−2

−1

01

2

φ12

φ 21

Figure 2-11. 21φ , 22φ Space when eigenvalues are 0.2 and 0.8 respectively.

2.9 Conclusion

We have proposed a method for monitoring multivariate processes based on two

charts, a common cause plot based on predicted values from a fitted VARIMA model and

a special cause chart that is a Hotelling’s T2 chart based on the residuals from that model.

We provided a detailed discussion of a two-dimensional VAR(1) model, and showed the

performance of the proposed method relative to a traditional T2 chart applied directly to

the unfiltered data. In this case, no uniformly superior performance can be claimed; in

many cases the performance depends on the time series parameters and the shift. It

should be anticipated that when the time series model gets more complicated, and there

are more parameters, it will be more complicated. However, our discussion of the

behavior for a VAR(1) model provides enough evidence to indicate potential benefits of

the new method.

We like to point out that the ARL performance is not the only reason we think the

residuals based method should be used. The fundamental issue is that rather than

Page 66: Topics in Multivariate Time Series Analysis: Statistical

48

proposing an ad hoc algorithm, we ought to identify and estimate a time series model

before we attempt to control it. Without knowing the autocorrelation structure and fitting

an appropriate time series model to a process, we do not fully understand the process.

That said, we do not suggest replacing any of the traditional control charts. The

combination of both will help achieve better control of processes. The advantage of the

model based method is that it takes into account the auto- and cross- correlation.

We agree with Box and Paniagua-Quiñones (2007) that it makes more sense to

model industrial process as weakly non-stationary process in which case the traditional

Hotelling’s 2T would be inappropriate. Since the residuals are approximately random if

the model fits well, all of the assumptions of traditional multivariate process control

charts are met. Moreover, the residual chart can be used to detect changes in the structure

of the time series while observed data based methods are not necessarily sensitive to

structural changes.

From the perspective of applications, the problem with the proposed method is the

difficulty of fitting VARIMA models. However, for a cross-correlated and autocorrelated

multivariate process, we find it often worthwhile to understand the process better by

fitting a model. Even if we only implement Hotelling’s T2 chart based on raw data, we

still need to know the model to control the ARL0, as discussed in Section 2.5. Besides the

difficulty of fitting VARIMA models there is also the problem that the number of

parameters increase rapidly with an increase of the time series dimensions. A process

with more than ten dimensions is almost intractable. Thus we suggest considering various

dimension reducing approaches, as will be discussed in Chapter 4 of this dissertation.

Page 67: Topics in Multivariate Time Series Analysis: Statistical

49

CHAPTER 3

A CLASS OF MARKOV CHAIN MODELS FOR AVERAGE RUN LENG TH COMPUTATION FOR AUTOCORRELATED PROCESSES

The ARL of conventional control charts is typically computed assuming temporal

independence. However, this assumption is frequently violated in practical applications.

Alternative ARL computations have often been conducted via time consuming and not

necessarily very accurate simulations. In this chapter, we develop a class of Markov chain

models for evaluating the run length performance of traditional control charts for

autocorrelated processes. We show extensions from the simplest univariate AR(1) model

to the general univariate ARMA(p, q) model, and the multivariate VARMA(p, q) time

series case. The results of the proposed method are compared with simulation results. The

Markov chain methods for processes discussed in this chapter are implemented in

MATLAB and are available online as supplemental materials.

3.1 Introduction

As introduced in Section 2.1.4, the ARL has become the standard metric for the

evaluation of the performance of control charts. In the literature it is often assumed that

the observations are temporally independent. This typically implies that the control chart

statistic being charted (monitored) also is independent. If independent, the run length has

a geometric distribution with mean equal to 1−p , where p is the probability of the

statistic falling outside the control limits. However, in many important applications, the

data and hence the statistic monitored is autocorrelated, which implies that the traditional

ARL computations may be seriously misleading.

Page 68: Topics in Multivariate Time Series Analysis: Statistical

50

Autocorrelation of the monitored statistic occur sometimes even when the data is

assumed independent. For example the Cumulative Sum (CUSUM) and the

Exponentially Weighted Moving Average (EWMA) are autocorrelated even when the

underlying process being monitored is independent. To compute the ARL for such

situations, Brook and Evans (1972) introduced a Markov chain approximation and used

that to analyze the performance of the univariate CUSUM control chart. Lucas and

Saccucci (1990) applied the same technique to provide design recommendations to the

univariate EWMA control chart. Runger and Prabhu (1996) extended this idea to a

multivariate EWMA control chart.

To deal with an autocorrelated process, ad-hoc control charts as well as model

based monitoring algorithms have been developed. For example, Berthouex, Hunter and

Pallesen (1978) suggested filtering the process data through an autoregressive integrated

moving average (ARIMA) time series model and applying regular control charts to the

residuals. Alwan and Roberts (1988) extended this idea and proposed a special-cause

chart based on residuals from fitting a univariate time series model as well as a common

cause chart based on the one-step prediction. Jiang (2001) used the Markov chain

approximation to compute the ARL for his ARMA chart, a special case of which includes

the special-cause chart (Alwan and Roberts, 1988) as showed in Jiang et al. (2000) for

univariate processes.

That said, quality practitioners frequently, advertently or inadvertently, apply

conventional control charts such as Shewhart control charts and Hotelling’s T2 control

charts when processes are autocorrelated. Moreover, for robustness studies it is important

to compute the ARL performance under alternative assumptions. The purpose of this

Page 69: Topics in Multivariate Time Series Analysis: Statistical

51

chapter is to provide a class of Markov chain approximations for ARL computations

when a given time series model of the process is assumed known. We hope that a more

accurate ARL computation will provide guidance for setting more appropriate control

limits for conventional control charts. We acknowledge that conventional control charts

sometimes are applied to monitor non-stationary processes, with no constant first and

second moments. However, we confine our discussion to stationary time series processes.

The novelties are that we work on conventional control charts, that we extend the Markov

chain approach from univariate cases to multivariate cases, and that we use polar

coordinates to divide the in-control state space in the multivariate case.

This chapter is organized as follows. In Section 3.2 we review the basic idea of

using a Markov chain approximation to calculate the ARL illustrated with an example of

a univariate AR(1) model. In Section 3.3 we present an extension to the general

ARMA(p,q) model. In Section 3.4 we extend the Markov chain approximation to the

multivariate time series case and illustrate it with an example of a first order Vector

Autoregressive VAR(1) model. Computational accuracy, timing and strategies are

discussed in Section 3.5. Finally we present a conclusion and discussion in Section 3.6.

3.2 The Univariate AR(1) process

Consider an AR(1) model 1t t tz z aφ −= +% % where ta is white noise with 0}{ =taE ,

{ } 1tV a = , and t t tz z µ= −% , ( )t tE zµ = . We further assume that the parameters of the

AR(1) model are known. When the mean is constant, we can assume without loss of

generality that 0tµ = and t tz z=% . We declare observations outside the interval

],[ ccC −= to be out-of-control. The Markov chain method for computing the ARL is

Page 70: Topics in Multivariate Time Series Analysis: Statistical

52

based on the Markov property of the AR(1) model. As illustrated in Figure 3-1, the real

line is divided into 1s+ subintervals, with s subintervals within the interval C . One

straightforward way of dividing C is [ 2 ( 1) / , 2 / ), 1, 2, ,iS c c i s c ci s i s= − + − − + = K and

then to define the ( )1 -ths+ “interval” as 1 ( , ) [ , )sS c c+ = −∞ − ∪ −∞ , out-of-control zone.

We then define a Markov chain tX with states }1,,2,1{ +sΚ and let

( )1

1

s

t t iiX i I z S

+

== ⋅ ∈∑ , where ( )I ⋅ is indicator function defined such that tX is in state i

if t iz S∈ . The ARL computation then proceeds as follows: Starting from any in-control

state, how many steps does it take to arrive at the out-of-control state for the first time.

Symbolically we denote this time by { }1 0min 1, s.t. 1| 1nn X s X sτ = ≥ = + ≠ + .

Figure 3-1. Markov Chain State Space for AR(1) model.

Brook and Evans (1972) proved that ARL is given by

( )1τ ′=-1

φ I - R 1 , (3.1)

where φ is a 1s× dimensional vector where i-th element ( )iφ is the probability

( )0 0| 1P X i X s= ≠ + , 1,2,...,i s= . Further R is a s s× matrix with ( ),i j -th element

( )1|ij t tr P X j X i−= = = , (3.2)

and 1 is a 1s× unit vector.

Brook and Evans (1972) used factorial moments to prove (3.1). However, we

provide an alternative proof in APPENDIX C.

Consider a step mean shift for the AR(1) model at time T such that

Page 71: Topics in Multivariate Time Series Analysis: Statistical

53

( )1 1t t t t tz z aµ φ µ− −− = − + , (3.3)

where,

( )~ 0,1

0

.

t

t

a N

t T

t Tµ

δ<

= ≥

(3.4)

Combining (3.4) and (3.3) we get

( )

1

1

11 .

t t

t t t

t t

z a t T

z z a t T

z a t T

φδ φφ δ φ

+ <

= + + = − + + >

(3.5)

Now suppose the whole process starts from an in-control state at time 1t T= −

and that a step-shift in the mean occurs at time T. The probability that the shift is detected

at time T is

( ) ( )

( )( )

( ) ( )

( )

1 1 1 1 1

1 2 1 2 11 1 1

1 1

1| 1 |

,1

,

T T T s T s

c c

z aT s T s c cc

T s zc

p P X s X s P z S z S

f z f z z dz dzP z S z S

P z S f z dz

φ δ

− + − +

+ − + − −

− +−

= = + ≠ + = ∈ ∉

− −∈ ∉= = −

∉∫ ∫

(3.6)

where ( )zf ⋅ is the probability density function (pdf.) of tz and ( )af ⋅ is the pdf. of ta .

If the shift has not been detected at time T, which means 1TX s≠ + , the whole

Markov chain starts with initial probabilities,

( ) ( )( )

( )( )

( )( )

( )

1

1

1

1

1

1

1

| 1, 1

, 1, 1

1, 1

, 1| 1

1| 1

| 1

.1| 1

T T T

T T T

T T

T T T

T T

T T

T T

i P X i X s X s

P X i X s X s

P X s X s

P X i X s X s

P X s X s

P X i X s

P X s X s

φ −

= = ≠ + ≠ +

= ≠ + ≠ +=

≠ + ≠ +

= ≠ + ≠ +=

≠ + ≠ +

= ≠ +=

≠ + ≠ +

(3.7)

Page 72: Topics in Multivariate Time Series Analysis: Statistical

54

The nominator of )(iφ is given by

( )( ) ( )

( )

1 2 1 2 1

1| 1,

i

i

c u

z ac lT T c

zc

f z f z z dz dzP X i X s

f z dz

φ δ−

− −= ≠ + =

∫ ∫

and the denominator is

( ) ( )1 111| 1 | 1

s

T T T TiP X s X s P X i X s− −=

≠ + ≠ + = ≠ ≠ +∑ .

Moreover the transition probability after time T is

( )( )

( )

( ) ( )( )

( )

1

11

1 2 1 2 1

,|

1

.

i j

i j

i

i

t j t i

ij t tt i

u u

z al l

u

zl

P z S z Sr P X j X i

P z S

f z f z z dz dz

f z dz

φ φ δ

−−

∈ ∈= = = =

− − −=∫ ∫

(3.8)

Using (3.1), we can compute the ARL as

( )( )

( ) ( )( )1 1 1

1 1

1 1 1

1 1 1 ,

ARL p p

p p

τ= ⋅ + − +

′= ⋅ + − +-1

φ I - R 1 (3.9)

where 1p , φ and R are computed through (3.6), (3.7) and (3.8) respectively.

To validate this Markov chain method we computed results for selected

parameters of three different AR(1) models for both the Markov chain method and the

simulation method as showed in Table 3-1. The standard deviation of each simulated

ARL was estimated on the entire sample of simulation outputs and is recorded after the

simulated ARL entries (± s.d.). Undoubtedly, the more the simulation rounds, the smaller

the standard deviation of the simulated ARL. The simulations were each run 90,000 times

to give a reasonable standard deviation; so were the simulations in the remaining part of

this chapter. For the Markov chain method, the number of subintervals, s, was chosen to

Page 73: Topics in Multivariate Time Series Analysis: Statistical

55

be 10. Both methods have been implemented in MATLAB and executed at a DELL

personal workstation (Intel Pentium Dual CPU E2180@2GHz, 2GHz, 1.99G of RAM).

Computing time has been recorded and provided in parenthesis following the ARL entries.

Notably, for the same underlying time series model, computing time for the Markov

chain method is independent of the size of the shift and the parameters while the

simulation time is nearly proportional to the ARL, i.e., the larger the ARL, the longer the

simulation time. As it can be seen, the results from the Markov chain method match the

simulation results very well while the Markov chain method took significantly less time

than the simulation method.

Table 3-1: Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on AR(1) models.

0φ = 0.5φ = 0.6φ = − MC Simulated MC Simulated MC Simulated 0δ = 300.00

(0.78s) 299.51 ± 1.00

(315.09s)

322.50 (0.83s)

323.52 ± 1.08

(323.52s)

341.81 (0.81s)

342.67 ± 1.14

(360.88s)

0.5 zδ σ= 129.24 (0.78s)

128.76 ± 0.43

(136.53s)

147.82 (0.81s)

148.15 ± 0.49

(156.45s)

138.57 (0.81s)

139.07 ± 0.46

(147.22s)

zδ σ= 37.70 (0.75s)

37.87 ± 0.12

(40.92s)

47.06 (0.78s)

47.36 ± 0.16

(51.23s)

40.08 (0.81s)

40.16 ± 0.13

(43.67s)

3.3 Extension to the Univariate ARMA(p, q) process

The Markov chain method can easily be applied to the AR(1) model because the

process tz is Markovian. For more general ARMA(p, q) models, we need to use a

transformation before the Markov chain method can be applied; see e.g. Box, Jenkins and

Reinsel (1995). We provide a few illustrations.

Page 74: Topics in Multivariate Time Series Analysis: Statistical

56

3.3.1 Extension to the Univariate AR(2) Process

Consider an AR(2) model 1 1 2 2t t t tz z z aφ φ− −= + +% % % , where ta is white noise with

0}{ =taE , { } 1tV a = , and t t tz z µ= −% , ( )t tE zµ = . Again for a process with constant

mean, we may without loss of generality assume that 0tµ = and t tz z=% . This univariate

AR(2) model can be presented in a multivariate AR(1) (Markovian state space) model as

11 2

1 2

1

1 0 0 .t t

tt t

z za

z z

φ φ −

− −

= +

(3.10)

The new statistic [ ]1t t tz z−′=w is Markovian. Thus the methodology from

Section 3.2 immediately applies. The state space for tw is illustrated in Figure 3-2.

Although tz and 1tz − are correlated, the control mechanism is imposed on observations at

each single time point. Thus the in-control state space is a square which is subdivided

into an m m× grid, each sub-square representing one state and all the points outside the

in-control square the ( )1 -thm m× + state. Notice that the transition matrix is

( ) ( )2 21 1m m+ × + but very sparse. The transition probability from state [ ]1j i to

[ ]2k j is only non-zero when 1 2j j= since “ tz ” at this time point has to become “1tz − ”

at the next time point.

Page 75: Topics in Multivariate Time Series Analysis: Statistical

57

Figure 3-2. Markov Chain State Space for AR(2) model.

Again consider a step shift model

( ) ( )1 1 1 2 2 2t t t t t t tz z z aµ φ µ φ µ− − − −− = − + − + , (3.11)

with ta and tµ follow (3.4) as well. A four-stage model for tz follows immediately:

( )

( )

1 1 2 2

1 1 2 2

1 1 1 2 2

1 2 1 1 2 2

1 1

1 1.

t t t

t t tt

t t t

t t t

z z a t T

z z a t Tz

z z a t T

z z a t T

φ φδ φ φφ δ φ φ

φ φ δ φ φ

− −

− −

− −

− −

+ + < + + + =

= − + + + = + − − + + + > +

(3.12)

Assume 1p is the probability that the shift is detected right away (t T= ), 2p is

the probability that the shift is not detected until the second step ( 1t T= + ), and tz

follows a Markov chain model for 1t T> + . The ARL is then

( ) ( )( )( )1 1 2 1 2 11 1 2 1 1 2ARL p p p p p τ= ⋅ + − ⋅ + − − + . (3.13)

Comparisons between the Markov chain method and simulations for the AR(2)

process is provided in Table 3-2. In the Markov chain method, m was chosen to be 10 and

thus s was 100. Again, the results from the Markov chain method closely match those

from the simulations but took significantly less time.

Page 76: Topics in Multivariate Time Series Analysis: Statistical

58

Table 3-2. Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on AR(2) models.

1

2

0

0

φ

φ

=

= 1

2

0.5

0.2

φ

φ

=

= 1

2

0.5

0.2

φ

φ

= −

= −

MC Simulated MC Simulated MC Simulated

0δ = 300.71 (9.03s)

301.09 ± 1.00

(368.82s)

369.88 (8.61s)

368.15 ± 1.22

(449.95s)

311.93 (8.67s)

312.47 ± 1.04

(381.72s)

0.5 zδ σ= 129.36 (8.41s)

129.71 ± 0.43

(160.00s)

181.29 (8.67s)

181.40 ± 0.60

(222.95s)

129.86 (8.59s)

130.28 ± 0.44

(160.22s)

zδ σ= 37.71 (8.51s)

37.55 ± 0.12

(47.98s)

62.51 (8.59s)

62.40 ± 0.21

(78.08s)

36.78 (8.62s)

36.99 ± 0.12

(47.18s)

3.3.2 Extension to the Univariate ARMA(1, 1) Process

For the ARMA(1, 1) model 1 1t t t tz z a aφ θ− −= + −% % with the assumptions as above,

the state space vector [ ]t t tz a ′=w can be transformed to be Markovian process as

follows

1

1

1

0 0 1 .t t

tt t

z za

a a

φ θ −

− = +

(3.14)

The state space for tw is illustrated in Figure 3-3. Theoretically control limits are

only imposed on tz , not on ta , implying that the in-control state space will be an infinite

area. However, as a pragmatic device we imposed a large artificial limit ]6,6[− on ta so

that the in-control state space becomes a finite rectangle. It has to be noted that by doing

this the computed ARL will be shorter than it really is because an abnormal ta will be

Page 77: Topics in Multivariate Time Series Analysis: Statistical

59

considered to be out-control without upsetting tz . However the bias is small with control

limits on ta chosen to be large.

Figure 3-3. Markov Chain State Space for ARMA(1,1) model.

The rectangle is divided into an m n× grid with each sub-square representing one

state and all the points outside the in-control rectangle representing the ( )1 -thm n× +

state. The transition matrix is ( ) ( )1 1mn mn+ × + but is again very sparse. In APPENDIX

D we show that once the initial state is fixed, two parallel lines with slopes of 1 will be

determined and that the ending states can only be between the two lines or by crossed

through the lines.

For illustration we consider a step shift ARMA(1, 1) model. Thus a three-stage

model for tz is

( )

1 1

1 1

1 11 .

t t t

t t t t

t t t

z a a t T

z z a a t T

z a a t T

φ θδ φ θφ δ φ θ

− −

− −

− −

+ − <

= + + − = − + + >

(3.15)

Similarly, the ARL is given by

( )( )1 1 11 1 1ARL p p τ= ⋅ + − + , (3.16)

Page 78: Topics in Multivariate Time Series Analysis: Statistical

60

with 1p again being the probability that the shift is detected at time T and 1τ determined

by Markov chain model after time 1T + .

Comparison between the Markov chain method and the simulations for the

ARMA(1,1) is showed in Table 3-3. For the Markov chain method, both m and n were

chosen to be 15 and thus s was 225. Again, the results from the Markov chain method

closely match those from the simulations except for a slight discrepancy in case 2

( 0.8, 0.5φ θ= = ). For ARMA(1,1) model, the two methods are comparable in computing

time. The realization of the Markov chain method on ARMA(1,1) process requires a set

of control flow statements that drastically increase the computing time compared with

cases for AR(1) and AR(2) models. MATLAB is especially inefficient with control flow

statements; had it been implemented with C/C++ or other programming languages, the

increased computing time might not be as significant.

Table 3-3. Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on ARMA(1,1) models.

0.5

0.8

φθ=

=

0.8

0.5

φθ=

=

0.4

0.2

φθ= −

=

MC Simulated MC Simulated MC Simulated

0δ = 303.85

(238.65s)

301.49 ± 1.01

(361.39s)

316.63 (242.43s)

322.13 ± 1.07

(385.47s)

326.60 (201.39s)

326.80 ± 1.09

(389.91s)

0.5 zδ σ= 128.47 (238.00s)

128.12 ± 0.42

(154.23s)

146.00 (247.62s)

153.21 ± 0.51

(183.69s)

133.56 (201.92s)

133.07 ± 0.44

(159.72s)

zδ σ= 36.62 (228.26s)

36.05 ± 0.12

(45.28s)

47.00 (244.73)

51.14 ± 0.17

(62.69s)

38.25 (192.14s)

38.30 ± 0.13

(48.14s)

Page 79: Topics in Multivariate Time Series Analysis: Statistical

61

3.3.3 Extension to the Univariate ARMA(p, q) Process

For the ARMA(p, q) model 1 1 1 1... ...t t p t p t t q t qz z z a a aφ φ θ θ− − − −= + + + − − −% % % , a

similar extension can easily be made by constructing a Markovian state space vector. A

general model for ARMA(p, q) model can be written in state space format as,

11 2 1

1 2

1

1

1

1

1 0 0 0

0

0 1 0 0 0

0 1

0 0 0 1 0

t tp q

t t

t p t p

t t

t q t q

z z

z z

z z

a a

a a

φ φ φ θ θ −

− −

− + −

− + −

− − = +

L L L

L L L L L

M MO O O O O O M M

O O O O

M O O O O O O

M MM O O O O O O M

M MM O O O O O O M

L L L

0

0 .

ta

M

(3.17)

For any ARMA(p, q) model, (3.17) provides a p q+ dimensional state space

vector model. An alternative classical state space model, see e.g. Reinsel (1997), for the

ARMA(p, q) model is given by

11

11 2 1

0 1 0 01

0 0 1 0

0 0 0 1,

t t t

rr r r

ψε

ψφ φ φ φ

−− −

= +

Z Z

L

L

M M M M MM

L

L

(3.18)

where ( ) ( ) ( )ˆ ˆ ˆ0 , 1 , , 1t t t tz z z r′ = − Z L , ( )ˆtz j is the expected value of t jz +

conditional on all the observations up to time point t , iφ ’s are the autoregressive

coefficients from the ARMA(p, q) model ( 0iφ = if i p> ) and jψ ’s are the coefficients

for the ARMA model converted to infinite MA form, 0t j t jj

z ψ ε∞

−==∑ . Note that the

dimension r of tZ , which is computed as ( )max , 1p q+ , is smaller than that of model

Page 80: Topics in Multivariate Time Series Analysis: Statistical

62

(3.17). Note further that the first element of tZ , ( )ˆ 0tz , is essentially tz , on which we

will impose the control limit for the computation.

We acknowledge that as either p or q increases, the complexity of the

computations increase geometrically. However, for most practical applications applying

the principle of parsimonious model building, p and q rarely exceed 2.

3.4 Extension to the Multivariate ARMA(p, q) process

In many industrial applications the processes to be monitored are multivariate. An

extension from univariate time series model to multivariate time series model is therefore

desirable. We demonstrate this extension with the example of the VAR(1) model of

dimension k . To facilitate easy visualization, we will choose 2k = .

Consider a VAR(1) model 1t t t−= +z Φz a% % where ( )tE =a 0, ( )tVar =a I and

again without loss of generality, t t t= −z z µ% where t =µ 0 . The state space for tz is

shown in Figure 3-4. The in-control area is an ellipse. Using polar coordinates it can be

divided into an m n× mesh of sectors each representing one state and the large sector

outside the in-control ellipse representing the ( )1 -thm n× + state. The transition matrix is

a ( ) ( )1 1mn mn+ × + matrix. By subdividing the in-control zone after converting to polar

coordinates rather than using a Cartesian grid we achieve better accuracy. The polar

coordinate transformation can easily be generalized to higher dimensions.

Page 81: Topics in Multivariate Time Series Analysis: Statistical

63

Figure 3-4. Markov Chain State Space of VAR(1) model.

In the multivariate case we use a step shift model similar to that is used for the

univariate case. The result is again a three-stage model for tz given as

( )

1

1

1 .

t t

t t t

t t

t T

t T

t T

+ <

= + + = − + + >

Φz a

z δ Φz a

I Φ δ Φz a

(3.19)

Equation (3.19) looks similar to (3.5) and works almost the same way. The

difference is that computing the elements of transition matrix is more tedious:

( )( )

( )

( ) ( ) ( )( )

( )( )

1

11

2 2 1 1 2 2 1 1

2

,

,|

, ,

,,

u u u ui i j j

l l l li i j j

u ui i

l li i

t j t i

ij t tt i

r r

ar r

r

zr

P z S z Sr P X j X i

P z S

f g r g r J d dr d dr

f g r J d dr

γ γ

γ γ

γ

γ

γ γ γ γ

γ γ

−−

∈ ∈= = = =

− − −=∫ ∫ ∫ ∫

∫ ∫ δ

I Φ δ Φ (3.20)

where the ( ) , g ⋅ ⋅ is the transformation from polar to Cartesian coordinates and J is the

corresponding Jacobian matrix.

Comparisons between the Markov chain method and simulations for the VAR(1)

are shown in Table 3-4. Note that unlike the univariate case, the step shift δ is a two

dimensional vector. Thus we need two polar parameters, α the angle and λ the length,

Page 82: Topics in Multivariate Time Series Analysis: Statistical

64

to describe it. For the Markov chain method, m and n were chosen to be 10 and 4

respectively, and thus s was 40. Again, the results from the Markov chain method match

the results from the simulations very well. However, both methods took significantly

more computing time than their counterparts in univariate cases.

Table 3-4. Comparison of computed ARL between the Markov chain method (MC) and the simulation method (Simulated) on VAR(1) models.

0 0

0 0

Φ =

0.2 0.4

0.1 0.1

Φ =

0.4 0.3

0.5 0.1

− Φ = −

MC Simulated MC Simulated MC Simulated

0

0

λα=

= 299.93

(668.35s)

299.98 ± 1.00

(999.69s)

305.24 (690.81s)

302.92 ± 1.01

(1053.70s)

322.66 (680.50s)

325.97 ± 1.09

(1119.50s)

0.5

0

λα=

= 108.45

(665.01s)

108.71 ± 0.36

(360.20s)

112.59 (697.92s)

112.21 ± 0.37

(406.50s)

114.22 (672.64s)

116.07 ± 0.39

(399.67s)

1

0

λα=

= 57.21

(675.75s)

57.17 ± 0.19

(190.55)

60.23 (691.02s)

60.12 ± 0.20

(217.94s)

60.23 (682.42s)

60.49 ± 0.21

(213.59s)

The extension to a VARMA(p, q) process is similar to the one going from the

univariate AR(1) to the univariate ARMA(p, q) process. However, the computation

becomes more complicated, especially when the vector dimension is larger than two.

3.5 Computation Efficiency

The Markov chain approximation that we have proposed raises a number of

computational issues that we will briefly discuss in this section.

Page 83: Topics in Multivariate Time Series Analysis: Statistical

65

3.5.1 How Fine Should the Mesh be?

To compute the ARLs, we need to determine the number s of subintervals into

which the in-control region is divided. ARLs have been computed against a set of s

values and are illustrated in Figure 3-5 through Figure 3-8. For comparison we also

simulated the ARLs. However, note that a simulated ARL is a pseudo random number.

Hence we provide for the simulations the ARL plus the standard deviation ( ..dsARL± )

to judge the convergence of the Markov chain method.

For the AR(1) model, s was chosen to be 10 for the computation in Table 3-1.

However, for the AR(1) case 1 ( 0φ = ), Figure 3-5 shows that 2=s provides result

accurate enough; for AR(1) case 2 (0.5φ = ), Figure 3-6 shows that 4=s works well

enough. Thus we infer that the higher the autocorrelation, the larger s has to be to achieve

the same level of accuracy. In fact, when the autocorrelation is 0, there is no need for the

Markov chain method; instead the ARL can be computed via the geometric distribution.

Page 84: Topics in Multivariate Time Series Analysis: Statistical

66

s

AR

L

20151050

300.5

300.0

299.5

299.0

298.5

299.51

298.51

300.51

Figure 3-5. ARL v.s. s for AR(1) model, 0φ = .

s

AR

L

302520151050

325

324

323

322

321

320

323.52

322.44

324.6

Figure 3-6. ARL v.s. s for AR(1) model, 0.5φ = .

Page 85: Topics in Multivariate Time Series Analysis: Statistical

67

For the AR(2) model, s was chosen to be 100 for the computation in Table 3-2.

However, Figure 3-7 shows for case 2 (1 20.5, 0.2φ φ= = ), 36=s or even 16=s provides

sufficient accuracy.

s

AR

L

4003002001000

370.0

367.5

365.0

362.5

360.0

368.15

369.37

366.93

Figure 3-7. ARL v.s. s for AR(2) model, 1 20.5, 0.2φ φ= = .

For the AR(1,1) model, s was chosen to be 225 for computation of Case 2

( 0.8, 0.5φ θ= = ) in Table 3-3. Figure 3-8 does show slower convergence compared with

Figure 3-5 through Figure 3-7. However, 81=s or 144=s provides sufficient accuracy.

Page 86: Topics in Multivariate Time Series Analysis: Statistical

68

s

AR

L

6005004003002001000

324

322

320

318

316

314

312

310

322.13

323.2

321.06

Figure 3-8. ARL v.s. s for ARMA(1,1) model, 0.8, 0.5φ θ= = .

As a rule of thumb, in univariate cases, we have found that for the one-dimension

in-control region 10=s , and for the two-dimension in-control region, 100=s provides

sufficient accuracy. For the two-dimension AR(1) case, we have found that 40=s works

well.

One way to improve the computation accuracy of the Markov chain method is

through assuming the functional relation ( ) ( ) 2/ /ARL s ARL A s B s= ∞ + + between the

ARL and s as suggested by Brook and Evans (1972); see also Jiang (2001). By

computing pairs of ARL and s, OLS regression provides a reasonable estimate of

( )ARL ∞ . For the four cases shown in Figure 3-5 through Figure 3-8, the ( )ARL ∞

estimates are 300.0, 322.9, 370.6, 319.8 respectively. This regression model appears to

improve the accuracy. However, the least square estimates of the regression coefficients A

Page 87: Topics in Multivariate Time Series Analysis: Statistical

69

and B vary from case to case. Thus there does not appear to be a universal set of

coefficients that apply to every condition to save computation cost.

3.5.2 Matrix Computation

As was pointed out above, the transition matrices for AR(2) and ARMA(1,1)

cases are sparse. Figure 3-9 illustrates the transition matrices graphically for these two

cases. In the AR(2) case, with 100=s , the non-zero entries are 1000 out of a 10000 total

entries or 10%. In the ARMA(1,1) case, with 225=s , the non-zero entries account for

8027 out of 50625 entries or about 16%. In MATLAB, the function “sparse” can be

invoked to reduce memory storage and accelerate computing time.

0 10 20 30 40 50 60 70 80 90 100

0

10

20

30

40

50

60

70

80

90

100

nz = 1000

(a) AR(2), 100s= (b) AR(1,1), 225s=

Figure 3-9. Sparse transition matrices.

In general, matrix computation may be subject to large computation errors,

especially for large dimension sparse matrices. However, a 100100× matrix for the AR(2)

process, or a 225225× matrix for the ARMA(1,1) process, is by today’s standards not a

large matrix. To measure the approximation error of a large sparse matrix A , the

computational inverse for ( )I - R with R defined in (3.2), we use the Frobenius norm of

the matrix ( )( )⋅I - A I - R . If A is close to ( ) 1−I - R , ( )⋅A I - R is close to I and the

Page 88: Topics in Multivariate Time Series Analysis: Statistical

70

Frobenius norm of the matrix ( )( )⋅I - A I - R will be close to 0. For the 100100× matrix

for AR(2) computation, the Frobenius norm is 3.4133e-014, which indicates the

approximation error is very small.

Moreover the matrix computation is not where the major computing time was

spent. In the AR(2) case with 100=s , for example, the integral computation using the

lattice method (Sloan and Joe, 1994) to calculate the non-zero elements of the transition

matrix took 9.92s while the matrix computation only took 0.02s.

To reduce the integral computing time, one alternative is to use midpoint

multiplication. However, in that case, a finer mesh is required to achieve similar accuracy

and a much larger transition matrix will be needed. In spite of the computing cost saved

for each entry in the transition matrix, the computation efficiency will be offset by the

larger number of entries to be computed, and a larger, sparser transition matrix which will

result not only costly computing, but also raises accuracy problem because of the size.

3.6 Conclusion and Discussion

In this chapter we have proposed a class of Markov chain models for the

approximate ARL calculations for control algorithms where the monitored processes

exhibit autocorrelation. The suggested Markov chain approximation is versatile and we

show that it is easily generalized to multivariate time series process. It also performs well

compared to simulations.

One potential issue is deciding how fine a mesh to divide the in-control space to

make the computation as accurate as needed. Clearly the further it is divided the better

the results. However, a linear increase in the number of mesh points and hence states

Page 89: Topics in Multivariate Time Series Analysis: Statistical

71

result in a quadratic increase in the dimension of the transition matrix and thus the

computational effort. Fortunately, if the numerical computations of the multiple integral

based on lattice method is used for the transition probabilities as demonstrated in this

chapter, the accuracy is fairly high even if the in-control space is divided into only 10

grid points for the univariate AR(1) model and a 10 4× grid for the VAR(1) model.

Further the CPU time was only seconds for the univariate AR(1) models and about 10

minutes for the two-dimensional VAR(1) model, which is considerably shorter than what

would be needed for comparable accuracy using simulation methods.

Page 90: Topics in Multivariate Time Series Analysis: Statistical

72

CHAPTER 4

COMMON FACTORS AND LINEAR RELATIONSHIPS OF MULTIVARIATE TIME SERIES

Due to the ever-growing data made available by modern technologies, the effort

to understand business processes is often challenged by co-existing high-dimensionality

and autocorrelation. High-dimensionality obscures the true latent factors. In traditional

multivariate literature, Principal Components Analysis is the standard tool for dimension

reduction. For autocorrelated processes, however, PCA fails to take into account the time

structure information, i.e., the autocorrelation information. Thus it is arguable that PCA is

still the best choice. In this chapter I propose an enhanced dimension reduction method

which by design takes into account both cross-correlation and autocorrelation information.

I demonstrate through both a case study and a simulation study that the proposed method

is better than PCA in dimension reduction and discovering latent factors for multivariate

autocorrelated processes.

4.1 Motivation

Figure 4-1 shows a plot of the Furnace data introduced in Section 1.1.1. One

interesting pattern is that all the dimensions of the data seem to move up and down at the

same time. A similar pattern can be found in the Hog data displayed in Figure 1-3. In fact,

it is common to observe such “co-movement” in multivariate time series processes.

Consider for example the stock market. Stock prices move up and down as random walks.

However, there seems to be an invisible hand that keeps most of the stocks moving in the

same direction, the direction indicated by the Dow Jones Industrial Average index (DJIA)

or the Standard and Poor’s 500 index (S&P500).

Page 91: Topics in Multivariate Time Series Analysis: Statistical

73

Time

Tem

pera

ture

80726456484032241681

650

600

550

500

450

400

Variable

z3z4z5z6z7z8z9

z1z2

Figure 4-1. A traditional time series plot of temperatures measured at nine different locations of a large industrial furnace.

This phenomenon leads us to believe that the large dimension of multiple time

series often might be driven by a smaller number of common sources, often referred to as

common factors or latent variables. In the furnace example, a common factor could be an

overall trend computed by taking the average of the temperatures across the furnace. In

the Hog data example, it is conjectured that the GDP, a proxy for the overall economy,

could be a common factor. Similarly, in the stock market, the market factor, suggested by

Sharpe (1963, 1964) for the Capital Asset Pricing Model (CAPM), may serve as a

common factor.

To discover the common factors for a multivariate time series process is an

important issue from many perspectives. The immediate benefit of dimension reduction is

that it frees us from the so-called “curse of dimensionality.” It is often a formidable task

to fit a high-dimensional time series model with a large number of parameters. The

problem becomes more accessible once we can reduce the problem to focusing on only a

few common factors. Furthermore, finding a small number of common factors helps us

Page 92: Topics in Multivariate Time Series Analysis: Statistical

74

interpret the variability of the data. A reasonable short list of common factors may not

only explain most of the in-sample variability, but also a large portion of the out-of-

sample variability. A further reason is most relevant in the quality control context. Rather

than overwhelming our eyes with every single observation, we may only need to monitor

a few common factors. Last but not least, if a linear combination of the observed

variables does not contain any common factors, it implies that there might be a stationary

relationship among these variables. This relationship often provides insights into the

underlying structure of the data and hence allows us to interpret the data. In the

econometrics literature, the Error-Correction Model (ECM, see Engle and Granger, 1987)

is build upon relationships of this kind.

4.2 Literature Review

Principal Components Analysis, Factor Analysis, and Canonical Correlation

Analysis (CCA) have been used as tools to discover common factors in the general

multivariate setting. These techniques involve calculations based on the contemporary

variance-covariance matrix. Thus PCA in particular does not utilize any time-related

information. For multivariate time series, endeavors have been made to devise better

methods. Badcock et al (2004) devised a time-structure based but not a model based

method. Meanwhile other researchers start with assuming a certain model and base their

methods on that model. Approaches taken by them include the Canonical Correlation

Analysis by Box and Tiao (1977), the maximum likelihood estimation method for

cointegrated model by Johansen (1988), the Reduced Rank model by Velu, Reinsel and

Wichern (1986) and the Dynamic Factor model by Peña and Poncela (2004). In this

Page 93: Topics in Multivariate Time Series Analysis: Statistical

75

section, important literature is reviewed with an emphasis on the method by Box and Tiao

(1977).

4.2.1 The PCA Method

PCA applies to either the variance-covariance matrix zΣ or the correlation matrix

zρ of the observed data tz . I will focus on the covariance matrix based PCA only. The

method is based on finding a set of projection vectors ie which are mutually orthonormal

and sequentially maximize the variance of the transformed variables i t′e z (Johnson and

Wichern, 2002):

1 2 1, ,...,max

i i

i z i

i i−⊥

′e e e e

e Σ e

e e .

(4.1)

Unfortunately PCA is developed on the contemporary covariance matrix only.

Thus this method does not capture any structure in the data that manifests itself in the

autocorrelation and temporal cross-correlation structure. However, auto- and lagged

cross- correlation are key features of time series processes. Thus we question the

usefulness of the variability criteria used in PCA to decompose the data. Specifically we

question the ability of PCA to retrieve the common factors in time series data.

4.2.2 The Box-Tiao Method

Box and Tiao (1977) introduced a new projection method based on decomposing

the data according to “predictability” instead of “variability.” Their method, which we

will call the Box-Tiao method, is essentially a Canonical Correlation Analysis between

the vector of current observations and a set of past or lagged observations.

Page 94: Topics in Multivariate Time Series Analysis: Statistical

76

The derivation of the method is based on assuming a typical vector autoregressive

time series model

( )1ˆ 1t t t−= +z z ε, (4.2)

where ( )1ˆ 1t−z is a linear combination of past observations, representing the best

prediction of tz made at time 1t − . Accordingly for the variance, we have

ˆz z ε= +Σ Σ Σ . (4.3)

where ( )z tVar=Σ z , ( )( )ˆ 1ˆ 1z tVar −=Σ z and ( )tVarε ε=Σ .

By performing a linear transformation im on both sides of (4.2), we get

( )1ˆ 1i t i t i t−′ ′ ′= +m z m z m ε

. (4.4)

The Box-Tiao method then amounts to finding a set of transformation vectors im ,

ki ,,1Κ= , which sequentially maximize the predictability of the transformed variables

i t′m z . The predictability is defined as how much variability of the transformed variable is

explained by that of its best linear estimate at time 1t − . Defined in this way is it similar

to the 2R criterion used in regression. The problem is thus formulated as

ˆmax i z i

i z i

m Σ m

m Σ m .

(4.5)

If we set 1/2i i z′ ′=a m Σ , (4.5) becomes,

1/2 1/2 1/2 1/2 1/2 1/2

ˆ ˆ

1/2 1/2max maxi z z z z z i i z z z i

i z z i i i

− − − −′ ′⋅ ⋅ ⋅ ⋅=

′ ′⋅

m Σ Σ Σ Σ Σ m a Σ Σ Σ a

m Σ Σ m a a. (4.6)

Page 95: Topics in Multivariate Time Series Analysis: Statistical

77

It is now clear that the sequence of maxima in (4.6) is the set of eigenvalues iλ ’s

of 1/2 1/2ˆz z z

− −Σ Σ Σ , ranked from the highest to the lowest, with the ia ’s being the

corresponding eigenvectors. More explicitly substituting back 1/2i i z′ ′=a m Σ we get

1/2 1/2 1ˆ ˆz z z i i i z z i i iλ λ− − −⋅ = ⇒ ⋅ =Σ Σ Σ a a Σ Σ m m . (4.7)

Thus it is immediately seen that the vector im is an eigenvector of 1ˆz z

−Σ Σ ; the

corresponding eigenvalue iλ is the predictability of the i-th transformed variable.

The Box-Tiao components with predictability close to zero or close to one have

interesting implications. Those close to zero represent transformed time series that

basically are unpredictable random shocks. Thus these linear combinations of the original

time series are stable relationships among the component variables. They are related to

the concept of cointegration which we will pursue in more details in Section 4.3.2. The

Box-Tiao components with eigenvalues close to one represent nearly non-stationary time

series.

The original work by Box and Tiao (1977) is based on assuming a stationary

vector time series. For non-stationary multivariate time series, we may therefore question

the viability of the Box-Tiao method because it depends on the covariance matrices

which are not defined in this situation. Velu et al. (1987) justified the Box-Tiao method in

this case by analyzing asymptotic behaviors of the covariance matrices.

The Box-Tiao method provides an alternative to the PCA method and is more

appealing in many ways. The foremost reason perhaps resides in our intuition that

common factors differentiate themselves from each other in terms of their dynamics

Page 96: Topics in Multivariate Time Series Analysis: Statistical

78

rather than their variability, and it is the dynamics, not the variability, that does not

change over time.

4.2.3 PCA on ˆ aΣ

So far I have focused on the variance and autocorrelation of }{ tz . Another

interesting perspective is to look at the random shocks ta , which drive the time series in

multivariate ARIMA processes through tt BB aΘzΦ )()( = .

Tiao, Tsay and Wang (1993) illustrated the usefulness of several different linear

transformation methods by studying 1-month, 3-month and 6-month Taiwan’s interest-

rate. They first fitted a VAR(4) model and then applied PCA to the residual covariance

matrix ˆaΣ . For this data set they observed that ˆ

aΣ was nearly singular. It turned out that

by doing so, they were able to achieve parsimony parameterization and reveal interesting

and meaningful features of the interest-rate. They also showed that this approach

produced transformations very similar to the Box-Tiao transformation.

4.2.4 The Badcock Method

Badcock et al. (2004) developed a method for retrieving “temporally structured”

components. Specifically they were seeking components that exhibit distinct differences

in the persistence of autocorrelation. Their method was based on earlier work by Switzer

(1985) used in image reconstruction.

Above we defined ( ),l t t lCov −=Γ z z . After applying a linear transformation if to

tz , the autocovariance of the transformed variable i t′f z is i l i

′f Γ f . The Badcock method

Page 97: Topics in Multivariate Time Series Analysis: Statistical

79

then amounts to finding a set of transformation vectors if , ki ,,1Κ= , which sequentially

maximize the autocorrelation of the transformed variables i t′f z :

max.

i l i

i z i

′Γ

f f

f Σ f (4.8)

This idea has some similarities to the Box-Tiao method but is essentially different.

I will discuss this method in more detail in Section 4.3.3.

4.2.5 Dynamic Factor Model

In a dynamic factor model it is assumed that the observed vector of k time series

are generated by kr < common factors ty , plus a measurement error:

( ) ( ) ,

t t t

y t y tB B

= +

=

z Py ε

Φ y Θ a (4.9)

where tε is a white-noise sequence with full-rank covariance matrix εΣ . The matrix P is

a k r× loading matrix and ty is assumed to follow a r-dimensional VARMA(p,q) process.

In what follows ty represents the common factors. If we further assume that ( )y BΦ and

( )y BΘ are diagonal then the model is called an uncoupled-factors model. Otherwise it is

called a coupled-factors model.

A key problem in dynamic factor modeling is to identify the number r of latent

factors. In the stationary case Peña and Box (1987) developed a procedure to identify this

number by looking at the eigenvalues of lagged covariance matrices. Peña et al. (2004)

generalized this method to the non-stationary case by analyzing the asymptotic behavior

of re-defined covariance matrices.

Page 98: Topics in Multivariate Time Series Analysis: Statistical

80

4.3 Reflections on the Literature

4.3.1 The Box-Tiao Method and Canonical Correlation Analysis

Canonical correlation analysis originally developed by Hotelling (1936) seeks to

identify and quantify the associations between two sets of variables. It focuses on finding

linear combinations of the variables in one set and linear combinations of another set of

variables that sequentially have the highest correlation. The idea is to first determine the

pair of linear combinations having the largest correlation. Next, the pair of linear

combinations having the largest correlation among all pairs uncorrelated with the initially

selected pair, etc. The pairs of linear combinations are called the canonical variables, and

their correlations are called canonical correlations. For technical details, see Johnson and

Wichern (2002).

I now briefly show that for a VAR(p) process, the Box-Tiao method is the

canonical correlation analysis between one set tz and another set { }1 2, , ,t t t p− − −z z zK . The

predictability of the i-th Box-Tiao component corresponds to the square of the canonical

correlation of the the i-th pair of canonical variables.

The VAR(p) model is readily expressed as,

1 1t t p t p t− −= + + +z Φ z Φ z aL . (4.10)

The Yule-Walker equations (see Reinsel 1997) for this model implies that

1 10 1 1

1 0 22 2

1 2 0 .

p

p

p pp p

− −

− −

′ ′ ′ ′ = ′ ′

Γ ΦΓ Γ Γ

Γ Γ ΓΓ Φ

Γ Γ ΓΓ Φ

L

L

M M O MM ML

(4.11)

Page 99: Topics in Multivariate Time Series Analysis: Statistical

81

The Canonical Correlation analysis between the set { }tz (set 1) and the set

{ }1 2, , ,t t t p− − −z z zK (set 2) amounts to finding the eigenvectors of 1 111 12 22 21− −Σ Σ Σ Σ , where

iiΣ is the variance-covariance matrix of set i, and ijΣ is the variance between set i and set

j. Specifically with the Yule-Walker equations (4.11) substituted in, 1 111 12 22 21− −Σ Σ Σ Σ can be

expressed as,

110 1 1

1 0 21 21 2

1 2 0

10 1 1

1 0 21 21 2

1 2 0

, , ,

, , ,

.

p

p

z p

p pp

p

p

z p

p pp

− −−

− −

− −−

− −

′ ′ ′

′ ′ = ′

ΓΓ Γ Γ

Γ Γ Γ ΓΣ Γ Γ Γ

Γ Γ ΓΓ

ΦΓ Γ Γ

Γ Γ Γ ΦΣ Φ Φ Φ

Γ Γ ΓΦ

L

LK

M M O M ML

L

LK

M M O M ML

(4.12)

As demonstrated in section 4.2.2, the Box-Tiao transformations are the

eigenvectors of 1ˆz z

−Σ Σ , which is the same as (4.12) by recognizing that

( )ˆ 1 1

10 1 1

1 0 2 21 2

1 2 0

, , ,

.

z t p t p

p

p

p

p pp

Cov − −

− −

− −

= + +

′ ′ = ′

Σ Φ z Φ z

ΦΓ Γ Γ

Γ Γ Γ ΦΦ Φ Φ

Γ Γ ΓΦ

L

L

LK

M M O M ML

(4.13)

Thus we see the relationship between Box-Tiao transformation and the canonical

correlation analysis. The first Box-Tiao component is the linear transformation of the set

{ }tz which has the largest correlation with a corresponding linear transformation on the

second set { }1 2, , ,t t t p− − −z z zK .

Page 100: Topics in Multivariate Time Series Analysis: Statistical

82

Note that the predictability of the i-th Box-Tiao component corresponds to the i-th

eigenvalue iλ of (4.12), while the correlation of the the i-th pair of canonical variables

corresponds to the square-root of the i-th eigenvalue iλ of (4.12). Note also that

eigenvalues essentially are the 2R ’s in regression analysis.

4.3.2 The Box-Tiao Method, Cointegration and Co-movement

When a univariate time series have d unit-roots, it is said to be integrated of order

d, denoted as I(d). For non-stationary multivariate time series, it is possible for there to be

a linear combination of integrated variables that is stationary. These variables are said to

be cointegrated. More specifically, the components of the vector [ ]1 2, ,t t t ktz z z ′=z K are

said to be cointegrated of order d, b, denoted by ( )~ ,t CI d bz if all components of tz are

integrated of order d and there exists a vector [ ]1 2, , kβ β β ′=β K such that the linear

combination t′β z is integrated of order d b− where 0b > . See Engle and Granger (1987)

for more technical details.

If cointegration does exist, being able to recognize it may help us avoid over-

differencing the data and prevent the fitted model from being non-invertible. In the

econometric literature, cointegration is an important issue because it implies a long-run

economic equilibrium.

Developed a decade earlier than the cointegration econometrics literature, the

Box-Tiao method is based on the same idea. Consider an integrated multivariate time

series process. If there exists a Box-Tiao component with predictability near zero, then

the related Box-Tiao component is a purely random noise series which implies a stable

Page 101: Topics in Multivariate Time Series Analysis: Statistical

83

relationship among the integrated component variables. Such a relation is generally called

co-movement. In the special case where the original vector time series is non-stationary

then it is called a cointegrated relationship. Note if a Box-Tiao component of non-

stationary vector time series has predictability between zero and one (i.e., the Box-Tiao

component follow a general ARMA model), it also indicates co-movement but not

cointegration. On the other hand, a Box-Tiao component with predictability near one, is a

latent variable that when mixed into all the component variables makes every component

variable look non-stationary. Table 4-1 summarizes how predictability λ and unit roots

determine whether the case is cointegration or co-movement.

Table 4-1. Cointegration and Co-movement.

Whether there are Unit Roots Yes No

0λ = Cointegration

Or Co-movement Co-movement

0 1λ< < Cointegration Neither

We conclude that the Box-Tiao method can be used to identify co-movement as

well as cointegration. Consider for example a vector time series where there is Box-Tiao

component with predictability equal to 0.6 as well as a component with eigenvalue close

to zero. Although the time series may not be integrated, the Box-Tiao component with

predictability 0.6 contributes to every component variables in the process and influences

them to co-move to some extent.

4.3.3 The Box-Tiao Method and the Badcock Method

The Box-Tiao method and the Badcock method have certain similarities. They are

both based on an assumption that common factors differentiate themselves from each

Page 102: Topics in Multivariate Time Series Analysis: Statistical

84

other in terms of time structure. The former uses predictability as the criterion while the

latter uses autocorrelation. In many cases, the two methods produce similar

transformations. For example, if the time series follows a VAR(1) model, the Box-Tiao

transformation vectors are eigenvectors of 1z z− ′Σ ΦΣ Φ while the Badcock transformation

vectors are of the eigenvectors of ( )1z z z− ′ ′+Σ ΦΣ Σ Φ .

Based on my research I find that the Box-Tiao method is more appealing. The

predictability criterion used by the Box-Tiao method is better because it takes into

account the whole history of tz while the Badcock method only considers the

autocorrelation at one particular lag. Badcock et al. (2004) suggests to pick the specific

lag in a more or less ad hoc and arbitrary fashion. Secondly, Badcock et al. (2004)

suggested using ( ) / 2l l−+Γ Γ to approximate lΓ without accessing the difference and its

impact. From this perspective, it appears to be an ad hoc method not based on the

particular time series model.

Despite these misgivings, the Badcock method is easier to use because it does not

necessitate the fitting of a model. However, this “advantage” may not be worth it as we

give up the benefits that come from knowing the underlying model.

4.3.4 A Simple Augmented Model

To gain better understanding of how the PCA and Box-Tiao methods work, I

propose a simple augmented model to study the performance of the two methods. The

model is a dynamic factor model with time-structured noise. It is a special case of the

Dynamic Factor Model (4.9).

Page 103: Topics in Multivariate Time Series Analysis: Statistical

85

In real world applications the noise is not necessarily “white.” For example, in the

furnace example, the ambient temperature could affect the temperature readings from the

thermocouples and it itself could be autocorrelated. Although this added noise has as

strong time structures as common factors, we assume that it has a small loading in the

factor model. When retrieving common factors, we are not particularly interested in this

added noise.

I will assume that the dimension of the observations tz is k, that there are kr <

dimensional common factors vector ty and s dimensional time structured noise factors

vector tn . Further I assume that k r s≥ + and the model is

[ ]

( ) ( )( ) ( ) ,

t t t t t

y t y t

n t n t

B B

B B

= +

=

=

t

t

yz = Py + Qn +ε P Q ε

n

Φ y Θ a

Φ n Θ b

M

(4.14)

where tε , ta and tb are white noises uncorrelated with each other. Further ( )y BΦ ,

( )y BΘ , ( )n BΦ , ( )n BΘ , aΣ and bΣ are diagonal matrices so that all time-structured

components and the noise are uncoupled. It is also assumed that tn has a smaller loading

than ty , i.e., ( ) ( )1~ 1~

max max , ij ijj s j r

q p i= =

<< ∀ , where ijp is the ij -th element of matrix P and

ijq is the ij -th element of matrix Q . ( )k r s× + matrix [ ]P QM is of full rank r s+ .

Without loss of generality, it is further assumed that y =Σ I and n =Σ I . The goal of

dimension reduction is to retrieve ty .

Page 104: Topics in Multivariate Time Series Analysis: Statistical

86

4.3.5 A Simulation Study

To gain insights to how PCA and the BT method works I now provide a

discussion of a simulation study based on (4.14). I choose r to be 2 and s to be 1. Three

independent time series were simulated. They were designed to differ in both variability

and autocorrelation structure. The model was

( )

1 1 1 1 1

2 2 1 2 2

1 1 1 1 1

0 0 5 0 where ~ ,

0 0.9 0 1

0.95 where ~ 0,0.05 .

t t t t

t t t t

t t t t

y y a aN

y y a a

n n b b N

= +

= +

0 (4.15)

The first series ty1 is white noise, but it has large variability. The second series

ty2 and the noise series 1tn are similar in time structure, both positively autocorrelated,

yet are different in variability. The variance of ty2 , denoted by 2

2yσ , is 11.11 and the

variance of 1tn , denoted by 1

2nσ , is 0.51. Figure 4-2 shows the time series plot of 1y , 2y

and 1n against the same scale. Obviously 1n is dominated by ty1 and ty2 in variability

while exhibiting a clear time series structure.

Page 105: Topics in Multivariate Time Series Analysis: Statistical

87

Index

2

0

-2

100806040201

3

2

1

0

-1

100806040201

2

1

0

-1

-2

y1 y2

n1

Figure 4-2. Time Series plot of 1y , 2y and 1n .

Next I create nine time series by adding up linear combinations of the underlying

common and noise factors [ ]t t tn ′′=x y M and a random noise vector tε

( )

1 1 11

2 2 22 9

19 9 9

2 9 1

6 3 2

4 1 1

9 7 2

where ~ ,4 4 1

6 9 3.

2 5 1

8 3 2

5 7 2

t t tt

t t tt

tt t t

zy

zy N

nz

ε εε ε

ε ε

= +

0 IM M M

(4.16)

Note that if not for tε the variance-covariance matrix of tz would be singular.

This added-in noise disguises the true dimensionality. Also note the linear transforming

matrix is chosen such that the weight on the noise factor 1n will not make it dominate the

Page 106: Topics in Multivariate Time Series Analysis: Statistical

88

common factors. This design assures 1n does not contribute much variability to the

observed data set tz .

The PCA method suggests the first two PCs explain 99.7% of variability.

However, when we compare them with 1y and 2y , as shown in Figure 4-3, the results is

not satisfying. The PCs are not able to track 1y and 2y accurately. Both are confounded

mix of underlying components.

Page 107: Topics in Multivariate Time Series Analysis: Statistical

89

Index

2

0

-2

9060301

3

2

1

0

-1

9060301

2

1

0

-1

-2

9060301

10

0

-10

-20

-30

80

40

0

-40

-80

y1 y2 n1

pca2 pca1

(a)

y1

2

0

-2

100-10-20-30

y2

2

0

-2

pca1

n1

80400-40-80

2

0

-2

pca2

(b)

Figure 4-3. (a) The Time Series plot of 1y , 2y , 1n and the first two PCs; (b) The Matrix

plot of 1y , 2y and 1n against two PCs.

The Box-Tiao Method provides interesting insights. Shown in Figure 4-4, the first

Box-Tiao component tracks 2y well while 1y is not tracked out at all.

Page 108: Topics in Multivariate Time Series Analysis: Statistical

90

Index

3

2

1

0

-1

9060301

2

0

-2

2

1

0

-1

-2

9060301

3

2

1

0

-1

2

1

0

-1

-2

9060301

2

1

0

-1

-2

y2 y1 n1

bt1 bt2 bt3

(a)

y1

2

0

-2

20-2

y2

2

0

-2

bt1

n1

20-2

2

0

-2

bt2 bt3

20-2

(b)

Figure 4-4. (a) The Time Series plot of 1y , 2y , 1n and the first three Box-Tiao

components; (b) The Matrix plot of 1y , 2y and 1n against the first three Box-Tiao

components.

This simulation result is not surprising. By design the Box-Tiao transformation

seeks the components with predictability ranking from the highest to the lowest. The

second common factor 2y has a clear time structure and a considerable variability. The

noise factor 1n has a clear time structure as well but because of its low variability, it

Page 109: Topics in Multivariate Time Series Analysis: Statistical

91

easily gets contaminated. The first common factor 1y does not have predictability at all

because it is white noise; it is thus hard to be singled out from other white noise.

No conclusion is attempted to draw from one simulation study, but it is

conjectured when either method is applied alone, problem happens. They are each

designed for a special function; either function working alone will not produce the true

underlying components.

4.3.6 Other Application of the Box-Tiao Method

Much of the past research built upon the Box-Tiao method focused on unit roots

and cointegration. Bossaerts (1988) suggested that the Box-Tiao components could be

directly tested for the presence of unit roots using the methods and critical values

provided by Dickey and Fuller (1979). Bewley et al. (1994) found that the Box-Tiao

estimator of the cointegrating parameter (the Box-Tiao eigenvectors) performs well

compared to Johansen’s (1988) maximum likelihood estimator. Bewley and Yang (1995)

proposed cointegration tests based on the Box-Tiao method and found that compared to

Johansen’s (1988) and Engle and Yoo’s (1987) tests, the Box-Tiao based tests are more

robust to correlation between the disturbances in the cointegrating relationships and those

generating the common trends. In a more recent paper by Anderson et al. (2006), it was

shown that compared with Johansen’s test (1988), the Box-Tiao based test shows promise

in dealing with small sample size, fat tails, heteroskedasticity and skewness.

Page 110: Topics in Multivariate Time Series Analysis: Statistical

92

4.3.7 Spurious Box-Tiao Components

Box and Tiao (1977) cautioned that their method has a potential problem if the

residual covariance matrix aΣ is singular. If aΣ is of rank 1k , it is possible to transform

the VAR(p) time series model of tz to

( ) ( )

( ) ( )1 111 12 1

12 221 22,

l lpt t tl

l llt t

B=

= +

z zΦ Φ b

z z 0Φ Φ (4.17)

where 1tb is of dimension 1k . In another words, the ( )1k k− -dimensional vector 2tz is

completely predictable from past values ( )1 t l−z and ( )2 t l−z , 1,2, ,l p= K . Thus it is

possible that eigenvalues close to one could come from either integrated components or a

singular aΣ . So far we have no method for distinguishing between these two cases.

Another singularity problem occurs when there is lagged linear redundancy in tz

(e.g., , , 1i t j tz z −= ). In this case, one of the canonical correlations would be exactly equal to

1. However, it does not mean that we have an integrated component at all. Rather it

comes from the lagged relationship and we want to know how to treat this case.

If we rewrite the VARMA model as ( ) ( )1t tB B−=z Φ Θ a% , then it is clear that that

singularity of either of the matrices ( )BΦ and ( )BΘ will cause singularity of the

observed data. As will be demonstrated with the furnace data in the next section, these

spurious components make it tricky to use the Box-Tiao method to tackle dimension

reduction problem alone.

Page 111: Topics in Multivariate Time Series Analysis: Statistical

93

4.4 A Combined method of both the PCA and the Box-Tiao Methods

The analysis in Section 4.3.5 suggests neither the PCA method nor the Box-Tiao

method on its own can retrieve common factors exactly. In Section 4.4.1 we further

demonstrate the problem of applying the Box-Tiao method individually through both the

furnace data and a simulated example. In Section 4.4.2 we introduce a two-step procedure

of sequentially applying the PCA method and the Box-Tiao method. In Section 4.4.3 we

prove the effectiveness of the combined method based on some approximation. In Section

4.4.4 we further demonstrate its effectiveness with a revisit to the simulated example

discussed in Section 4.3.5 and in Section 4.4.5 we revisit the furnace example.

4.4.1 The Problem of Using the Box-Tiao method Individually

Recall the furnace example we introduced in Section 1.1.1. Figure 1-2 shows

temperature measurements at nine locations of the ceramic furnace. By eyeballing each

individual series in Figure 1-2 we find none of them resemble an I.I.D process. They all

exhibit certain dynamic and even some non-stationarity. More interestingly, 6tz through

9tz all look alike; they reach peaks and troughs around the same time. It seems plausible

that the real dimension is much smaller than nine, say, two or three.

If we rely on the Box-Tiao method to pick the important factors, however, we

soon run into troubles. Assuming an AR(1) model, the Box-Tiao analysis generates a set

of components with predictability shown in Figure 4-5.

Page 112: Topics in Multivariate Time Series Analysis: Statistical

94

Factors Index987654321

1.0

0.8

0.6

0.4

0.2

0.0

Predictability of Box-Tiao Factors

Figure 4-5. Predictability of Box-Tiao Factors of the Furnace data.

Among the nine Box-Tiao components, five have predictability higher than 0.5

and eight of them are obviously autocorrelated. However, does it mean five or eight of

these components are important underlying factors? A plot of the correlation of these

Box-Tiao factors (labeled BT1~BT9) with each observed variable 1 9~z z , as shown in

Figure 4-6, tells an interesting story. The fourth Box-Tiao component BT4, for example,

despite of having a predictability as high as 0.71, has low correlation with all observed

variables (the highest correlation is with 3z and it is 0.45); or stated in another way, it has

limited contribution to the overall variability. Same argument applies to the second, fifth,

seventh, eighth and ninth components.

Page 113: Topics in Multivariate Time Series Analysis: Statistical

95

Z

1.0

0.5

0.0

1050

0

0

1.0

0.5

0.0

1050

1.0

0.5

0.0

1050

0

BT1 BT2 BT3

BT4 BT5 BT6

BT7 BT8 BT9

Correlation of BT1, BT2, BT3, BT4, BT5, BT6, BT7, BT8, BT9 vs Z

Figure 4-6. Correlation of each Box-Tiao factor with each dimension of observed data.

Together with our prior knowledge from Section 4.3.7 on how possible suspicious

components could emerge, we see the Box-Tiao method alone may not lead to

meaningful dimension reduction. The component with high predictability could result

from a highly correlated error term, lagged observation or could be a time-structured

noise factor. With what follows I further demonstrate this idea with a simulation.

The Box-Tiao analysis on the Furnace data reveals that there are highly

predictable components with limited contribution to variability. This confirms with the

idea of the proposed augmented model (4.14) where the noise factors are separated from

the common factors because of their little contribution to the overall variability. In the

following I illustrate with one realization of (4.14) how noise factors could affect the

Box-Tiao analysis. The observation tz and white noise tε are both 9 1× vectors. 9ε =Σ I .

3y =Σ I . 3n =Σ I . The loading matrix P and Q are also randomly generated and then

Page 114: Topics in Multivariate Time Series Analysis: Statistical

96

scaled to ensure ( )1~

max 1.5 , ij ijj r

p q i=

≥ ∀ . ty and tn are both simulated from a VAR(1)

model with first-order autoregressive coefficient matrix yΦ and nΦ respectively:

0 0 0 0.75 0 0

0 0 0 0 0.85 0

0 0 0.9 0 0 0.95y n

= =

Φ Φ

Only one common factor is autocorrelated. However, one run of this simulation

shows at least three retrieved Box-Tiao components are predictable with predictability

equal to 0.93, 0.47 and 0.26. The reason behind this is that the highly autocorrelated noise

factors contaminate the analysis.

4.4.2 A Two-step Combined Method

The PCA and Box-Tiao methods use contemporary and temporary covariance

information respectively and use different optimization criteria. However, both methods

consist of performing a linear transformation of the data. Hence if combined, the new

transformation will also be a linear transformation. It is not so easy to see what properties

a combined transformation will have. However, I conjecture a combination of both may

have interesting properties. The combined method consists of a two-step procedure:

Step 1: Apply the PCA method, keeping the first r components according to the

scree plot.

Step 2: Apply the Box-Tiao method on the remaining r components. The

transformed components will be the final results.

The idea behind this procedure is that in the first step, the noise, regardless of its

time structure, will be peeled off because dimensions with small variability will be

Page 115: Topics in Multivariate Time Series Analysis: Statistical

97

discarded. Thus after the first step, what is left over is hopefully a linearly transformed ty .

After applying the Box-Tiao method which will order linearly transformed variable

according to predictability, exactly the ty will appear as the final results.

Note that applying the two methods in a reverse order does not work. If we apply

the Box-Tiao method first, the variability of the transformed time series will be distorted

and it is no longer possible to order the time series according to variability as is hoped to

be done with PCA.

4.4.3 Demonstration of the Combined Method

In this section I demonstrate through deduction how the combined method

combines the merits of both the PCA and the Box-Tiao methods.

Proposition 1: PCA method cannot always retrieve the underlying components ty

from tz in the model (4.14).

Proof. I prove the Proposition 1 with an approximation that tε is 0. The same

approximation is also used in Proposition 2 and 3.

Since

[ ] tt

t

yz = P Q

nM ,

to be able to retrieve the first component, there has to be a transformation vector i′e , so

that [ ] [ ]1 0 ... 0i′ =e P QM , which means , 1i j j⊥ ∀ ≠e p . Similar pattern follows for

other transformation vectors.

Page 116: Topics in Multivariate Time Series Analysis: Statistical

98

Consider a simple case where 2m= , 2r = , 0s= and 1 2cos , 0≠p p . Further

assume that tε is very small and εΣ is close to 0 . In this two-dimension case, to retrieve

the first component, as implied above, 1 2⊥e p and through 1 2cos , 0≠p p we

infer 1 1cos , 1≠ ±e p , which means 1e is independent of 1p . To retrieve the second

component, 2 1⊥e p and 2 1⊥e e , which is impossible in two dimension space since 1e is

independent of 1p . This counterexample indicates that PCA is not capable to retrieve true

factors exactly.

In the meanwhile the Box-Tiao method alone cannot discriminate the noise tn

from the more important common factors ty . A heuristic argument will be given through

three steps.

Proposition 2: for any VARMA(p,q) process ( ) ( )t tL L=Φ x Θ ε , if all

, 1,...,j j p=Φ , , 1,...,j j q=Θ and εΣ are diagonal, then the Box-Tiao transformation of

tx is tx .

Proof: If jΦ , jΘ and εΣ are all diagonal, then ( ) ( )0 tCov=Γ x x is diagonal.

For the equivalent expression ( )-1ˆ 1t t t+x = x ε , define ( ) ( )( )0 -1ˆ ˆ 1tCov=Γ x x , thus we have

( ) ( )0 0 ˆ ε= +Γ x Γ x Σ . Since εΣ and ( )0Γ x are diagonal, then ( )0 ˆΓ x is also diagonal.

Now the Box-Tiao transformation vectors are the eigenvectors of ( ) ( )-10 0 ˆQ = Γ x Γ x .

Thus since ( )0Γ x and ( )0 ˆΓ x are both diagonal, Q is also diagonal. It then follows that

Page 117: Topics in Multivariate Time Series Analysis: Statistical

99

the eigenvectors of Q are the columns of I . Thus the Box-Tiao transformation of tx is

t t=Ix x .

Proposition 3: the Box-Tiao transformation is invariant to a full rank linear

transformation. Specifically if the Box-Tiao transformation of tx gives ty , the Box-Tiao

transformation on t t=z Tx is also ty , for any full rank square matrix T .

Proof: Suppose ( )-1ˆ 1t t t+x = x ε , ( ) ( )0 tCov=Γ x x and ( ) ( )( )0 -1ˆ ˆ 1tCov=Γ x x . We

then have ( )-1ˆ 1t t t t= = +y Tx Tx T ε , ( ) ( ) ( )0 0tCov ′Γ y = y = TΓ x T and

( ) ( )( ) ( )( ) ( )0 t-1 -1 0ˆ ˆ ˆ ˆ1 1tCov Cov ′=Γ y = y Tx = TΓ x T . Box-Tiao transformation vectors are

eigenvectors of ( ) ( )-10 0 ˆQ = Γ x Γ x for tx and eigenvectors of

( ) ( )-1 -10 0 ˆ ′ ′Q = Γ y Γ y = T QT% for ty .

If λ=Qm m , then -1 -1 -1λ′ ′ ′ ′=T QT T m T m . Now let TQTQ ′′= −1~ then

( ) ( )-1 -1λ′ ′=Q T m T m% . Further denote by -1′=m T m% , then m% is an eigenvector of Q%.

The Box-Tiao transformation of t t=y Tx associated with eigenvector m% is

1t t t

−′ ′ ′= =m y m T Tx m x% . Thus we see that this is equivalent to the Box-Tiao component

of tx associated with the eigenvector m of Q .

Finally based on Proposition 2 and 3 I demonstrate by way of a heuristic

argument that applying the Box-Tiao method alone retrieves both ty and tn in the

Page 118: Topics in Multivariate Time Series Analysis: Statistical

100

simplified model (4.14). I acknowledge that my argument is not a proof because it is

based on assuming that that tε is 0.

When m r s= + , as discussed above, the Box-Tiao transformation of [ ],t t′′ ′y n is

an identity transformation. Thus because of the invariance of the Box-Tiao transformation,

the linear transformation [ ][ ],t t t′′ ′z = P Q y nM is [ ],t t

′′ ′y n .

When m r s> + , the first r s+ components will still be [ ],t t′′ ′y n while the

remaining components will be 0.

My conclusion based on above reasoning is that the Box-Tiao transformation

retrieves both ty and tn . However, it cannot further distinguish ty from tn .

4.4.4 Simulated Study Revisit

As a demonstration of this two-step procedure, I try it on the simulated data from

Section 4.3.5. After the first step I kept the first two PCs and then transform those again

with the Box-Tiao method. Figure 4-7 shows how closely the two final retrieved

components match the true underlying components 1y and 2y . Component “pca12bt1” is

the first Box-Tiao component on the first two PCs; component “pca12bt2” is the second

Box-Tiao component on the first two PCs.

Page 119: Topics in Multivariate Time Series Analysis: Statistical

101

Index

3

2

1

0

-1

9060301

2

0

-2

9060301

2

1

0

-1

-2

9060301

2

1

0

-1

-2

2

1

0

-1

-2

y2 y1 n1

pca12bt1 pca12bt2

(a)

y1

2

0

-2

210-1-2

y2

2

0

-2

pca12bt1

n1

210-1-2

2

0

-2

pca12bt2

(b)

Figure 4-7. (a) The Time Series plot of 1y , 2y , 1n and the two PCA-BT components; (b)

The Matrix plot of 1y , 2y and 1n against the two PCA-BT components.

To see whether we can generalize the result above and justify the two-step

procedure for the combined method, a more sophisticated simulation was designed to

compare the three methods and see how well they retrieve the common factors.

The simulation is based on (4.14) and constructed as follows. The observation tz

and white noise tε are both 9 1× vectors. The variance-covariance matrix of tε is

Page 120: Topics in Multivariate Time Series Analysis: Statistical

102

( )9~ ,NεΣ 0 I . The underlying components ty and time-structured noises tn are both

3 1× vectors and their variance are ( )3~ ,y NΣ 0 I and ( )3~ ,n NΣ 0 I respectively. The

random variability contributed by tn and tε are at similarly low level, and much smaller

than the variability contributed by ty . The loading matrix P and Q are also randomly

generated and then scaled to ensure ( )1~

max 3 , ij ijj r

p q i=

≥ ∀ . ty and tn are both simulated

from a VAR(1) model with first-order autoregressive coefficient matrix yΦ and nΦ

respectively:

0 0 0 02 0 0

0 0.5 0 0 0.5 0

0 0 0.9 0 0 0.8y n

= =

Φ Φ

For each run, 100 observations of tz were generated. Three methods were applied

and for each of them, retrieved components were matched to the true factors and

correlation between each matched pair was calculated. Simulation was run 1000 times

and the averaged correlations were computed and shown in table 1.

Table 4-2. Correlation between each factor and its matching component.

Correlation Between PCA Box-Tiao Combined Method

1st factor and its matching component 0.7369354 0.2559151 0.960456 2nd factor and its matching component 0.7366181 0.7395445 0.9461788 3rd factor and its matching component 0.764179 0.966231 0.9831987

From Table 4-2 we see the combined method retrieves three components which

matches the true underlying components very well, with correlations close to one. The

Box-Tiao method is capable of matching the third factor 3ty because it has a obvious

time structure (1φ as high as 0.9). However, it does not retrieve the first factor 1ty at all.

Page 121: Topics in Multivariate Time Series Analysis: Statistical

103

This is because the first factor does not have a time structure and the Box-Tiao method

probably confounds the time-structured noise with the not-time-structured underlying

component. PCA works fine generally for all these factors but none of them has been

matched exactly. The correlations are all below 0.8.

Based on the simulations the combined method appears to be promising and it

provides intuition about how it works.

4.4.5 Furnace Data Revisited

Apply the two step scheme on the Furnance data: keep the first four PCs and then

conduct the Box-Tiao transformation. Figure 4-8 shows time series plots and

autocorrelation function plots of the four components.

Page 122: Topics in Multivariate Time Series Analysis: Statistical

104

1009080706050403020101

-4105

-4106

-4107

-4108

-4109

-4110

Time Series Plot of xpcabt1

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpcabt1(with 5% significance limits for the autocorrelations)

1009080706050403020101

537

536

535

534

533

Time Series Plot of xpcabt2

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpcabt2(with 5% significance limits for the autocorrelations)

1009080706050403020101

-1757

-1758

-1759

-1760

-1761

-1762

-1763

-1764

Time Series Plot of xpcabt3

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpcabt3(with 5% significance limits for the autocorrelations)

1009080706050403020101

-1279

-1280

-1281

-1282

-1283

-1284

Time Series Plot of xpcabt4

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpcabt4(with 5% significance limits for the autocorrelations)

Figure 4-8. The Four PCA-BT Factors on Furnace data: Time Series plots and Autocorrelation Fuction.

Page 123: Topics in Multivariate Time Series Analysis: Statistical

105

The last component has an interesting autocorrelation pattern. It seems that

autocorrelations at odd lags and those at even lags have totally separate tracks. It is as if

the last component is a channel for two different variables (possibly correlated) X and Y,

which are alternating to show through this channel. The two variables are autocorrelated

and cross-correlated. It coincides with the fact that there are two cooling devices

switching back and forth every 20 minutes to cool the furnace. Since this is hourly data,

two consecutive data records happen to be influenced by different cooling devices.

Further more, all four Box-Tiao components are confirmed to be important

common factors by their autocorrelation with the observed variables, as shown in Figure

4-9.

Z

0

86420

1.00

0.75

0.50

0.25

0.00

86420

1.00

0.75

0.50

0.25

0.00 0

PCABT1 PCABT2

PCABT3 PCABT4

Scatterplot of PCABT1, PCABT2, PCABT3, PCABT4 vs Z

Figure 4-9. Correlation between the four BT-PCA factors and each dimension of the observed data.

Page 124: Topics in Multivariate Time Series Analysis: Statistical

106

1009080706050403020101

3940

3939

3938

3937

3936

Time Series Plot of xpca1

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpca1(with 5% significance limits for the autocorrelations)

1009080706050403020101

2146

2145

2144

2143

2142

2141

2140

2139

2138

Time Series Plot of xpca2

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpca2(with 5% significance limits for the autocorrelations)

1009080706050403020101

191

190

189

188

187

Time Series Plot of xpca3

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpca3(with 5% significance limits for the autocorrelations)

1009080706050403020101

236.5

236.0

235.5

235.0

234.5

234.0

233.5

233.0

Time Series Plot of xpca4

Lag24222018161412108642

1.0

0.8

0.6

0.4

0.2

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

Autocorrelation Function for xpca4(with 5% significance limits for the autocorrelations)

Figure 4-10. The Four PCAs on Furnace data: Time Series plots and Autocorrelation Fuction.

Page 125: Topics in Multivariate Time Series Analysis: Statistical

107

The two-step scheme produces a satisfactory analysis. If, however, we stop at the

PCA step and keep the four PCs, what factors will we get? Figure 4-10 provides time

series plots and autocorrelation function plots of the first four PCs for comparisons.

We find out the alternating autocorelation effect appear in all last three PCs,

which indicates that they are all contaminated by the PCABT4 – the switching cooling

factor, to some extent. This further demonstrates how PCA treats data against the

variability criteria and may be incompetent to differentiate factors according to their

dynamics.

4.5 Business Applications

Dimension reduction has wide applications to many business fields. In this section,

I focus discussion on Quality Management and Finance.

4.5.1 Statistical Process Control

Dimension reduction methods, especially the PCA method has been suggested for

SPC; see Jackson (1991), Runger and Alt (1996). For a long time it was reasoned that

attention should be focused on principal components (PCs) with highest eigenvalues

(major PCs). However, it turns out that the PCs with the lowest eigenvalues (PCA

residuals) are also important. Runger and Alt (1996) suggested applying multivariate

control charts on the PCA residuals. His argument is that the major PCs reflect system

dynamics and common causes while the PCA residuals tend to be white noise and

channels whereby assignable causes may enter the system. Thus focusing on PCA

residuals better positions us in detecting assignable causes. This idea appears to be

analogous to the residual based Hotellings 2T chart I proposed in Chapter 2. The

Page 126: Topics in Multivariate Time Series Analysis: Statistical

108

difference is that time series model residuals are replaced by PCA residuals. Runger’s

methodology is innovative yet vulnerable in terms of the assumption that the PCA

residuals will be uncorrelated over time. This is not necessarily true. One counter

example was provided by Badcock (2004). A promising alternative is to use Box-Tiao

components with close to zero eigenvalues instead of PC components. I will refer to this

approach as Box-Tiao residuals. These components are close to white noise and most

possibly associated with assignable causes. Monitoring mechanism built on these

components will be succinct summary of the data yet provide an efficient tool to look

into assignable causes.

4.5.2 Applications in Asset Pricing

In this section I will only outline that dimension reduction also plays an important

role in Finance. Thus the method discussed herein in my thesis will apply quite widely.

However, it is not my goal to provide a lengthy discussion of problems in Finance.

An important area in Finance is Asset Pricing. In Asset Pricing, one key issue is

how to determine the market price for risk and appropriately measure risk for a single

asset assuming market equilibrium. One economic model used to solve this problem was

developed almost simultaneously by Sharpe (1963, 1964) and Treynor (1961), while

Mossin (1966), Lintner (1965, 1969) and Black (1972) developed it further. This model is

normally referred to as the Capital Asset Pricing Model (CAPM). In recognizing some

negative empirical results on testing CAPM and the strong assumptions CAPM requires,

a multifactor model was developed by Merton (1973), stating that any factors that

influence the growth of consumption should also price individual assets. Later the

arbitrage pricing theory (APT), formulated by Ross (1976), assuming that random returns

Page 127: Topics in Multivariate Time Series Analysis: Statistical

109

are linear functions of a number of economy-wide exogenous random factors and an

idiosyncratic random variable, offers a testable multifactor alternative to CAPM.

Although APT is more flexible than CAPM because CAPM is a special case of

APT, there is no guidance with respect to the number and the nature of the underlying

pricing factors. This not only creates a problem on testing APT, but also becomes a

problem itself. To address this, research was motivated to extract the true factors, testing

APT or a combination of both.

Typically, researchers relied on two approaches to identify the underlying factors.

One approach is to use a set of pre-specified macroeconomic variables/factors. Chen,

Roll and Ross (1980) showed that exposures to widely discussed macroeconomic

variables such as innovations in inflation and the term structure generate risk premium. A

more successful model was proposed by Fama and French (1992, 1993) and they, based

on previous literature, proposed using size and book-to-market equity as the other

underlying factors besides the overall market factor.

The other approach is to use statistical factor methods. Chamberlain and

Rothschild (1976), Lehmann and Modest (1988) used the PCA method to extract

underlying components. Connor and Korajczyk (1987) extended Chamberlain and

Rothschild’s work and proposed an asymptotic principal components technique. Roll and

Ross (1980) used maximum likelihood factor analysis and tested APT based on their

estimated factor loadings.

It would thus be interesting to replace the PCA method in the literature with the

Box-Tiao method, or even the combined method. The PCA method, though widely used,

does not necessarily provide satisfying results. It is not guaranteed that the Box-Tiao

Page 128: Topics in Multivariate Time Series Analysis: Statistical

110

method would work better, but it provides an interesting alternative to explore. To

compare, I apply both the PCA and the Box-Tiao methods to the return data on portfolios

formed on each size decile (stocks are ranked according to size – market capitalization,

and then divided into ten groups; portfolios are formed for each group). And then look at

the correlation between these empirical factors and theoretical proposed pricing factors –

the Market factor (Rm-Rf), the size factor (SMB) and the book-to-equity factor (HML).

Figure 4-11 provides a graphical illustration of such correlations. PCA1~PCA3 represent

the first three PCs. PCA12BT1 represents the Box-Tiao transformed first component on

first two PCs. PCA12BT2 is named in the same fashion. It is obvious that only PCA1 has

a reasonably high correlation with one of the theoretical factors – the Market factor while

the other two PCs do not. The two-step resulted factors are able to match both the Market

factor and the size factor; and the correlation is high (0.971 and 0.747 respectively). It is

obvious that Box-Tiao transformation refines the factors and improves the retrieving

results.

Page 129: Topics in Multivariate Time Series Analysis: Statistical

111

Rm-R

f20

0

-20

200-20 50-5

SMB

20

0

-20

pca1

HML

500-50

10

0

-10

pca2 pca3

100-10

pca12bt1 pca12bt2

50-5

Correlation between Rm-Rf, SMB, HML and pca1, pca2, pca3, pca12bt1, pca12bt2

Figure 4-11. How well do results from statistical factor retrieving methods match with theoretical proposed pricing factors.

Page 130: Topics in Multivariate Time Series Analysis: Statistical

112

CHAPTER 5

VISUALIZING PRINCIPAL COMPONENTS ANALYSIS FOR MULTIVARIATE PROCESS DATA

In this chapter, we suggest a simple method for visualizing the results of Principal

Components Analysis (PCA) intended to complement existing graphical methods for

multivariate time series data applicable for process analysis and control. The idea is to

visualize multivariate data as a surface that in turn can be decomposed with PCA. The

surface plots developed in this chapter are intended for statistical process analysis but

may also help visualize economic data and in particular co-integration.

5.1 Motivation

Data obtained from industrial processes are increasingly multivariate in nature.

Data visualization is critical to the understanding of complex data structures from such

processes and an important aid in their statistical analysis. The literature on data

visualization is highly developed for data structures with a single response. Cleveland

(1985, 1993) provide comprehensive overviews of methods and approaches while Tufte

(1983) and Wainer (2004) provide less technical, but lucid surveys of statistical graphics

and its history. Of course, anyone concerned about data analysis can benefit from reading

Tukey’s (1962) philosophical exposition of fundamental principles.

Graphical methods for multivariate data structures are less developed. The reasons

are obvious; multivariate data structures are more complex and constitute a richer class of

problems while we remain essentially confined to two-dimensional representations. A

reasonably comprehensive overview of the current state of the art can be gained by

consulting Gabriel (1985), Gnanadesikan (1997), Jolliffe (2002), Rencher (2002),

Johnson and Wichern (2002), Jackson (2003) and the references herein.

Page 131: Topics in Multivariate Time Series Analysis: Statistical

113

When it comes to visualization of multiple time series structures, the literature is

sparse. The standard references, Reinsel (1997), Lütkepohl (1993), Brockwell and Davis

(1991) and Peña, Tiao and Tsay (2001), hardly consider the topic except in terms of

standard univariate time series plots in separate panels or in overlaid format. An

exception is Cressie (1993) whose emphasis is on spatial structures rather than on time

series.

The purpose of this research is to describe a graphical data analytic method that

may help elucidate certain structural properties of multiple time series data, both

temporary and contemporary. As indicated, the traditional approach is to either overlay

the time series on a single plotting frame or plot the series on separate panels. Those

approaches provide insight into the temporary nature of time series, and in some cases,

cross correlations and other contemporary relationships. We do not suggest replacing

these useful approaches. Rather, the objective is to show that complementary insight

sometimes can be gained by also viewing multiple time series as a surface. Further, we

show how the surface can be decomposed in meaningful ways using PCA. One advantage

of the described method is that it can handle high-dimensional data vectors.

The application of our method is primarily geared toward statistical process

analysis and control where multiple time series are common. However, we have also

found that the method may be useful for multivariate time series data structures from

other domains.

The remainder of this chapter is organized as follows. In Section 5.2, we explain

the idea of using surface plots. In Section 5.3, we introduce a decomposition of the

original data set using principal components (PCs). In Section 5.4, we provide two

Page 132: Topics in Multivariate Time Series Analysis: Statistical

114

examples using the PCA techniques, one from statistical process control and one from

economics, both cases with obvious time series structure. We provide additional

discussion on approximating the surfaces and visualizing cointegration in Section 5.5 and

Section 5.6 respectively. Finally in Section 5.7, we provide a conclusion.

5.2 Graphical Representation of Multiple Time Series

Driven by the desire to visualize events and patterns over time, multiple time

series have, since the 10th or 11th century and certainly since the time of Playfair (1786),

typically been represented graphically as overlaid time series plots or as separate plots

carefully aligned to allow comparison; see Tufte (1983, p. 28). Most often, the plots have

a common abscissa (time) scale, but the ordinate may either have the same or different

scales for each time series. Figure 4-1 shows such a traditional plot of a multiple time

series from engineering process quality control. As was introduced in Chapter 4, this data

set contains simultaneous observations of the temperature at nine different locations of a

large industrial furnace. The temperatures are measured with thermocouples located at

equal distance along the length of the furnace. For technical reasons, the temperatures are

supposed to exhibit an arch-shaped profile, with the highest temperature (hot spot) at

location 4 (z4). It is important to the operation of the furnace to maintain this temperature

profile over time. This type of problem is common in industrial process monitoring and

control; see Kourti and MacGregor (1996).

Although Figure 4-1 conveys information about a common trend or comovement

along the time axis of the nine time series, it does not clearly provide an insight to the

contemporary relationship (the profile) among variables. In Figure 5-1, we have

represented the same data as a three-dimensional surface plot. The three axes are (i) a

Page 133: Topics in Multivariate Time Series Analysis: Statistical

115

dimension index, in this case, counting equidistant physical locations from one to nine; (ii)

time; and (iii) the data scale, in this case, temperature. From Figure 5-1, we see the

patterns of the nine time series evolve in time, as well as the interrelationship among the

variables. Figure 5-2 shows a contour plot of the same data. None of these representations

is “better” than any other. Rather, the three representations complement each other and

provide additional insight to the structure of the time series. Indeed, depending on the

orientation of the surface in Figure 5-1, some details may be hidden. However, an

additional advantage of plotting multivariate data as a surface is realized with modern

dynamic-graphics software, where the surface can be rotated and looked at from different

angles and perspectives, a feature that is difficult to convey in this paper presentation.

13

57

9

20

40

60

80

450

500

550

600

DimensionTime

Obs

Figure 5-1. Surface plot of the industrial furnace temperature profile.

Page 134: Topics in Multivariate Time Series Analysis: Statistical

116

Time

Lo

cati

on

8070605040302010

9

8

7

6

5

4

3

2

1

z

500 - 550550 - 600

> 600

< 450450 - 500

Figure 5-2. Contour plot of the furnace temperature profile.

5.3 Principal Components Decomposition

The idea behind PCA is to transform multivariate data by introducing a new set of

rotated orthogonal linear coordinates so that the sample variances of the transformed data

are in decreasing order of magnitude. A comprehensive introduction to PCA is provided

by Johnson and Wichern (2002). The application of PCA to time series data is described

by Tsay (2005), Tiao et al. (1993), and Peña et al. (2001). As pointed out by Anderson

(1963), when PCA is applied to time series data, one should be cautious especially for

small sample sizes because the sample variance may be seriously biased, or worse, if the

time series is nonstationary, not be particularly meaningful. In this context, we consider

PCA entirely as an exploratory, data-analytic, and descriptive tool.

Most applications of PCA involve an interpretation of the PCs as projections of

original data onto a new set of rotated coordinate axes. In this research, we project the

PCs back to the original coordinate axes. This allows us to decompose the original data

according to the PCs. With the aid of the surface plot, we then visualize what the PCs

Page 135: Topics in Multivariate Time Series Analysis: Statistical

117

represent in the original scales. In the following discussion, we have adopted the

notations used by Johnson and Wichern (2002); we refer to this text for definitions and

proofs.

Suppose we have p time series each of n observations. Let tkx denote the tth

observation of the kth time series. The pn× data set can then be summarized as the

matrix

=

npn

tpt

p

xx

xx

xx

Λ

ΜΜ

Ο

ΜΜ

Κ

1

1

111

X .

Further let S be the estimated dispersion matrix of X . We assume S is non-

singular. The ith principal axis is defined as the ith orthonormal eigenvector ˆie of S (we

use the “hat” circumflex on ie because ie is the sample eigenvector of the sample

dispersion matrix S, whereas ie is the eigenvector of the “true” dispersion matrix Σ ; the

same convention is used for other quantities below) and 1ˆ ˆ ˆ( , , )p=P e eK . Further

1

ˆ ˆˆ ˆ ˆ ˆp

i i ii

λ=

′ ′= =∑S PΛP e e where iλ are the eigenvalues of S. The ith principal component is

the projection of X on the ith principal axis, ˆˆ i iy = Xe . Now let 1ˆ ˆ ˆ( , , )p=Y y yK then

PXY ˆˆ = . Since 1ˆˆ −=′ PP , the inverse transformation is given by

11

1 1 1 1

ˆˆ ˆ ˆ ˆ ˆ ˆˆ ˆ ˆ ˆ, ,

ˆp p p p

p

′ ′ ′ ′ = = = = + + = + + ′

e

X YP YP y y y e y e C C

e

L M L L ,

Page 136: Topics in Multivariate Time Series Analysis: Statistical

118

where each of the pn× matrices ˆˆi i i′=C y e , 1, ,i p= K are of rank 1. Those p

rank-one matrices can be displayed as surface plots either individually or combined in

any way we please. Typically we may be interested in plotting 1

k

k ii=

=∑X C , pk ,,1Κ=

for an increasing sequence of k’s and hence essentially reconstruct or approximate the

original data set with a lower rank approximation leaving out only residual variability. As

we demonstrate in the following sections, this method of decomposition, reconstruction

and approximation of the data matrix X facilitates meaningful ways of looking at the

original data that provide visual images of PCs. A similar decomposition can be achieved

using singular value decomposition, see e.g. Seber (2004).

Above we explained the decomposition in terms of the eigenvector, eigenvalue

decomposition of the dispersion matrix. By replacing the dispersion matrix by the

estimated correlation matrix R , we can accomplish a similar decomposition. However,

because PCA is not scale invariant, those will typically not be the same. In some cases,

either one or the other decomposition will provide a more meaningful visualization.

We hasten to note that the above decomposition is based on the spectral

decomposition of a symmetric matrix, a purely algebraic result. Nowhere does it involve

distributional or temporal independence assumptions of the data. Thus, we consider it a

data analytic tool useful for data exploration and visualization but not for inference.

5.4 Examples

In this section, we use two examples to illustrate the idea of using surface plots

and PCA decomposition. The first example is the nine-dimensional furnace temperature

data introduced above and the second is a five dimensional data set from economics.

Page 137: Topics in Multivariate Time Series Analysis: Statistical

119

5.4.1 Example 1: The Furnace Temperature Data

As an illustration, we first apply the decomposition method to the furnace

temperature data. Table 5-1 provides a summary of a standard principal components

analysis based on the covariance matrix. Figure 5-3 shows the eigenvector elements

(loadings) versus the (vector) dimension indices.

Table 5-1. Principal Components Analysis Based on the Covariance Matrix of the Furnace-Temperature Data

Eigenvalue 251.22 18.24 4.19 2.13 0.83 0.40 0.10 0.04 0.01

Proportion 0.906 0.066 0.015 0.008 0.003 0.001 0.000 0.000 0.000

Cumulative 0.906 0.972 0.987 0.995 0.998 0.999 1.000 1.000 1.000

Eigenvectors 1e 2e 3e 4e 5e 6e 7e 8e 9e

1 0.088 -0.275 -0.279 -0.424 -0.157 0.763 -0.219 -0.073 0.011

2 0.244 -0.405 -0.411 -0.521 0.165 -0.526 0.163 0.069 -0.019

3 0.309 -0.603 0.123 0.466 -0.087 0.177 0.516 -0.049 -0.007

4 0.294 -0.364 0.188 0.226 -0.069 -0.186 -0.777 0.222 -0.040

5 0.258 0.032 0.790 -0.518 -0.132 0.000 0.145 0.039 0.004

6 0.397 0.116 0.041 0.048 0.504 0.029 -0.144 -0.695 0.257

7 0.403 0.270 -0.066 0.050 0.413 0.199 0.067 0.324 -0.662

8 0.422 0.292 -0.157 0.062 -0.023 0.094 0.091 0.499 0.664

9 0.432 0.302 -0.221 0.051 -0.702 -0.163 0.006 -0.315 -0.229

Page 138: Topics in Multivariate Time Series Analysis: Statistical

120

Dimension

0.4

0.2

0.0 0

8642

0.4

0.0

-0.4

0

0.5

0.0

-0.5

0

0.5

0.0

-0.5

0

0.5

0.0

-0.5

0

0.5

0.0

-0.5

0

8642

0.5

0.0

-0.5

0

0.5

0.0

-0.5

0

8642

0.5

0.0

-0.5

0

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 5-3. Plot of the Eigenvector Elements Versus Their Dimension Index Based on the Covariance Matrix of the Furnace Data with (a) for 1e , (b) for 2e and so forth.

From Table 5-1, we note that approximately 97% of the variability in the furnace-

temperature data is accounted for by the first two principal components. Further, we see

from Figure 5-3 that the first set of loadings (first eigenvector elements) indicate that the

first PC is an average across all temperatures. From Figure 5-3, we also see that the

second PC is a contrast between the front (locations 1, 2, 3, 4) and the back end (locations

5, 6, 7, 8, 9) of the furnace. Both of these interpretations make sense from an engineering

point of view. Figure 5-4 shows principal component surfaces for the first, second, third

and ninth components. Here we can take advantage of the surface plots to see how the

data profile gets decomposed. From Figure 5-4(a) and Figure 5-4(b), we see 1C and 2C

contribute substantially to the profile, while 9C , shown in Figure 5-4(d), does not

Page 139: Topics in Multivariate Time Series Analysis: Statistical

121

contribute much. In Figure 5-5, we have experimented with yet another representation of

the nine-dimensional time series, this time using a three-dimensional scatter plot.

1

3

5

7

9

20

40

60

80

0

200

400

600

DimensionTime

C1

13

57

920

4060

80

0

200

400

600

Dimension Time

C2

(a) (b)

13

57

920

4060

80

0

200

400

600

Dimension Time

C3

1

3

5

7

920

4060

80

0

200

400

600

Dimension

Time

C9

(c) (d)

Figure 5-4. Principal component surfaces for the Furnace Temperature Data: (a) surface of the first principal component 1C , (b) surface of the second principal component 2C , (c) surface of the third principal

component 3C , (d)surface of the ninth principal component 9C .

Page 140: Topics in Multivariate Time Series Analysis: Statistical

122

13

5

7

9

20

40

60

80

450

500

550

600

DimensionTime

Obs

Figure 5-5. A three-dimensional scatter plot of the furnace data.

In Figure 5-2, we already represented the nine dimensional time series as a

contour plot. The interesting aspect of this type of representation in the context of process

analysis and control is that, if the temperature profile was constant over time, the contour

lines would be parallel straight lines. In this case, we see that the process varied

significantly over time, especially for locations 7, 8 and 9. This was also visible in Figure

4-1 and Figure 5-1, but was perhaps made a bit more precise in Figure 5-2.

5.4.2 Example 2: The Hog Data

In Section 1.1.2, we introduced an econometrics example of so-called “hog data.”

Figure 1-3 shows the traditional graphical representation of the “hog data.” It is a five-

dimensional multiple time series consisting of 82 yearly observations from 1867 to 1948

of quantities relevant to the US hog trade. The five dimensions are hog supply, hog price,

corn price, corn supply, and farm wages, respectively. Please refer to Section 1.1.2 for

transformations applied to the original data.

Page 141: Topics in Multivariate Time Series Analysis: Statistical

123

Figure 5-6 is the surface plot of the standardized hog data. Here we may question

the ability of a two-dimensional plot to convey the information in a three-dimensional

surface, but again, with modern dynamic-graphics software, we can always rotate the

surface to see hidden details. Although the surface plot of the hog data does not have as

nice a profile as the temperature data, we include this example because once we apply the

decomposition on the data, the individual component surfaces provide us with intuition

about the data structure and cointegration, a topic we pursue in the Section 5.6.

Dim

ensi

on

1

2

3

4

5

Time [Year]

1880

1900

1920

1940

Standardized O

bs

−3

−2

−1

0

1

2

3

Figure 5-6. Surface plot of the Hog Data.

Because of the incommensurate scales of the five time series, we decided to base

the PCA on the correlation matrix. However, it could as well be carried out using the

dispersion matrix. Table 5-2 provides a summary of a standard PCA based on the

correlation matrix. Figure 5-7 shows the eigenvector elements (loadings) versus the

(vector) dimension indices.

Page 142: Topics in Multivariate Time Series Analysis: Statistical

124

Table 5-2. Principal Components Analysis based on the Correlation Matrix of the Hog Data

Eigenvalue 3.5870 1.0691 0.2534 0.0630 0.0275 Proportion 0.717 0.214 0.051 0.013 0.006 Cumulative 0.717 0.931 0.982 0.994 1.000 Eigenvectors 1e 2e 3e 4e 5e

1 0.446 -0.370 -0.730 -0.089 -0.351 2 0.491 0.227 0.513 -0.081 -0.661 3 0.386 0.631 -0.264 0.575 0.228 4 0.380 -0.635 0.358 0.494 0.283 5 0.516 0.100 0.076 -0.641 0.554

Dimension

0.48

0.36

0.24

0.12

0.00 0

531

0.50

0.25

0.00

-0.25

-0.50

0

531

0.6

0.3

0.0

-0.3

-0.6

0

531

0.50

0.25

0.00

-0.25

-0.50

0

0.50

0.25

0.00

-0.25

-0.50

0

(a) (b) (c)

(d) (e)

Figure 5-7. Plot of the eigenvector elements versus their dimension index based on the correlation matrix of the Hog data with (a) for 1e , (b) for 2e and so forth.

Next, we decompose the data in Figure 5-6 into five separate PC surfaces as

shown in Figure 5-8. We see from Table 5-2 that 1C , the first PC surface, representing

71.7% of the total variation and shown in Figure 5-8(a), can be interpreted as a general

trend visible in all five time series. We also notice the large increase around the years of

Page 143: Topics in Multivariate Time Series Analysis: Statistical

125

World War I, followed by a significant dip across all time series corresponding to the

effect of the 1930’s economic depression. Further, we see a rapid surge in general

economic activity at the advent of World War II.

Dim

ensi

on

1

2

3

4

5

Time [Year]

1880

1900

1920

1940

−3

−2

−1

0

1

2

3

C1

Dim

ensi

on

1

2

3

4

5

Time [Year]

1880

1900

1920

1940

−3

−2

−1

0

1

2

3

C2

(a) (b)

Dim

ensi

on

1

2

3

4

5

Time [Year]

1880

1900

1920

1940

−3

−2

−1

0

1

2

3

C3

Dim

ensi

on

1

2

3

4

5

Time [Year]

1880

1900

1920

1940

−3

−2

−1

0

1

2

3

C4

(c) (d)

Figure 5-8. Principal component surfaces for the Hog Data: (a) surface of the first principal component 1C ,

(b) surface of the second principal component 2C , (c) surface of the third principal component 3C ,

(d)surface of the fourth principal component 4C .

Page 144: Topics in Multivariate Time Series Analysis: Statistical

126

From Figure 5-7, we see that all elements of the first eigenvector (loadings) are

more or less the same and hence interpretable as an average across the time series. Thus,

we conclude that the first PC is a contemporaneous average that, over time, exhibits a

common trend. In other words, the first PC appears to be a general index of (agricultural)

economic activity. We believe that the visual image of this underlying trend as a surface

explains this well.

The second PC surface 2C shown in Figure 5-8(b), accounting for 21.4% of the

variability, seems to represent a noisy but stationary contrast between time series 2 and 3

versus 1 and 4, with less weight on time series 5. This pattern is further indicated by

inspection of the loadings of the second eigenvector shown in Figure 5-7. The third PC

surface, shown in Figure 5-8(c), represents 5.1% of the total variability. The primary

feature of this surface is that it is almost stationary in time, but shows a contrast between

time series 2 and 4 versus 1 and 3. The fourth PC surface 4C shown in Figure 5-8(d),

representing 1.3%, is essentially a non-informative random variation and may not be

worth serious interpretation. So is 5C , and, hence, it is not depicted here.

5.5 Further Discussion: Adding Up the Surfaces

We now proceed to plot the data by adding up each of the rank-one surfaces. In

other words, we suggest constructing rank 1, , 1j p= −K approximations of the original

data set. This is indicated in Figure 5-9 for the hog data showing , 2, 3k k =X . Note how

the surfaces approach the original surface as k increases. The surface plots show how

principal component surfaces add up to the original surface. By definition 1X is the same

Page 145: Topics in Multivariate Time Series Analysis: Statistical

127

as 1C in Figure 5-8(a). However, 2X , showed in Figure 5-9(a), is already a close

approximation to the original surface given in Figure 5-6. Adding the 3rd PC surfaces, as

showed in Figure 5-9(b), mostly adds details or noises and does not materially change the

general structure of the data. Indeed it might be easier to see the general structure of the

data from the surface plot of a lower rank approximation, in this case 2X , because most

of the noise is eliminated.

Dim

ensi

on

1

2

3

4

5

Time [Year]

1880

1900

1920

1940

−3

−2

−1

0

1

2

3

X2

Dim

ensi

on

1

2

3

4

5

Time [Year]

1880

1900

1920

1940

−3

−2

−1

0

1

2

3

X3

(a) (b)

Figure 5-9. Adding up the rank one principal component surfaces for the Hog data: (a) surface of the added-up first two principal components 2X , (b) surface of the added-up first three principal components 3X .

We have also added up the rank 1 surfaces for the furnace temperature data. Those

are shown in Figure 5-10, from 2X in (a) to 4X in (b), to 6X in (c), then to X in (d). It is

interesting to note that the first two PC’s explains 97.2% of the variability but the surface

plot reconstructed from only the first two PC’s (Figure 5-10a) is quite different from the

raw data surface (Figure 5-10d) in profile. The reason is the PC decomposition in this

Page 146: Topics in Multivariate Time Series Analysis: Statistical

128

case was based on the covariance; the first PC surface does account for the largest

variance component, but its location has been shifted and the profile reveals the shift.

1

3

5

7

9

20

40

60

80

0

200

400

600

DimensionTime

X2

1

3

5

7

9

20

40

60

80

0

200

400

600

DimensionTime

X4

(a) (b)

1

3

5

7

9

20

40

60

80

0

200

400

600

DimensionTime

X6

1

3

5

7

9

20

40

60

80

0

200

400

600

DimensionTime

Obs

(c) (d)

Figure 5-10. Adding up the rank one principal component surfaces for the Furnace data: (a) surface of the added-up first two principal components 2X , (b) surface of the added-up first four principal components 4X ,

(c) surface of the added-up first six principal components 6X , (b) surface of the added-up all principal

components X .

Page 147: Topics in Multivariate Time Series Analysis: Statistical

129

The above suggests two things: first, using PCA based on the covariance matrix

may show that the PCs explain most of the variance, but that does not imply that the data

reconstructed by the first PC will approximate the raw data. The first PC approximation

may still differ significantly in certain locations. Thus, instead of only looking at how

much variance the PCs explain, we may also want to consider how well the PCs

approximate the raw data in specific locations. Second, PCA reconstruction based on the

covariance is complex. One way to avoid this is to use the correlation matrix, in which

case the problem of location no longer is an issue. However, when working with the

correlation matrix, we essentially standardize the data and lose the profile information. In

the furnace example, where it was important to preserve the profile, we used the

covariance matrix; while in the hog-data example, we used the correlation matrix, and the

combined surface of the first two components approximated the raw-data surface quite

well.

5.6 Visualization of Comovement

Cointegration refers to the phenomenon occurring in econometrics where it often

is possible to find linear combinations of nonstationary time series that are themselves

stationary. More generally, we can sometime determine linear combinations of multiple

time series, stationary or nonstationary, that move together. This is called comovement;

see, e.g., Peña et al. (2001). Although it is not the main purpose here to discuss

cointegration or comovement, it is worth noting that the surface plots offer means for

visualizing comovement in a very intuitive fashion. For the hog data, we see that all the

five time series, when considered individually, appear to be nonstationary. However, if

we look at Figure 5-11, where the individual PCs are plotted as time series, we find the

Page 148: Topics in Multivariate Time Series Analysis: Statistical

130

first PC appears to be nonstationary. Indeed, it is integrated of order 1, I(1). In other

words, their first differences are stationary. If we now project the first PC back to the

original coordinates, as illustrated in Figure 5-8(a), we get five (p = 5) I(1) nonstationary

time series. Thus, the five time series appear to be linearly dependent or driven by a

single (one-dimensional) underlying trend. The general trend represented by the first PC

is an overall measure of the hog-related farm economy that drives all five time series.

Further, the last 4 PCs look stationary. This is also clearly visualized in Figure 5-8 and

Figure 5-11. This same phenomenon may occur for industrial processes. In the furnace

example, it is clear both from the physics of the process and from the plots that, when the

temperature in the furnace goes up or down, all nine temperatures more or less move up

or down concurrently. In other words, there might be individual differences from

thermocouple to thermocouple, but in general, the temperatures exhibit comovement.

Page 149: Topics in Multivariate Time Series Analysis: Statistical

131

Time [Year]

4

2

0

-2

-4

1939191518911867

2

0

-2

1939191518911867

1

0

-1

1939191518911867

0.6

0.3

0.0

-0.3

-0.6

0.4

0.2

0.0

-0.2

-0.4

PC1 PC2 PC3

PC4 PC5

Figure 5-11. Time series plots of the five principal components, PC1, …,PC5 for the Hog data.

5.7 Conclusion

Multivariate time series data occur in many areas of applications in engineering

analysis and quality control. Methods for visualizing patterns in such data, especially

when the dimensions are high, may help gain insight to the often complex structure of

such data. Methods currently available for visualizing multiple process time series are

relatively limited. The idea of using a surface plot makes it possible to visualize a

considerable number of time series in one single plot.

The decomposition of the surfaces via spectral decomposition of the covariance or

the correlation matrix appears to provide a useful way of dissecting a multivariate-

process data set into meaningful components. Indeed, we think that this way of

visualizing and interpreting PCs can be useful in many applications. Further, this

Page 150: Topics in Multivariate Time Series Analysis: Statistical

132

decomposition provides a simple way to explain the concept of co-integration and co-

movement.

Page 151: Topics in Multivariate Time Series Analysis: Statistical

133

CHAPTER 6

CONCLUSION

Due to the ever-growing data made available by modern technologies, the effort

to understand business processes is often challenged by co-existing autocorrelation and

high-dimensionality. My dissertation aims to address these two issues at the same time.

Operations Management, especially Quality Management, is one example of

fields in need of such a solution. Literature in designing multivariate control charts has

generally assumed observations I.I.D. while that in dealing with autocorrelation it largely

remains in one-dimension. In this dissertation, I fill this gap by offering a procedure to

treat autocorrelation and high-dimensionality simultaneously.

The procedure starts with understanding the process, which includes plotting the

data in different perspectives (Chapter 5), and retrieving meaningful common factors

(Chapter 4). When the observed dimension has been reduced to a handful of

representative common factors, it is suggested that a VARIMA time series model should

be fitted. Based on the model, a special cause chart on residuals and a common cause

chart on fitted values can then be run on Phase II data to effectively monitor the process

(Chapter 2). It is acknowledged that traditional Shewhart charts are still widely used

among practitioners despite the broadly existing autocorrelation. In such a situation, it is

urged to use a numerical computation method (Chapter 3) to calculate accurate ARL. The

calculation assuming an I.I.D. process can be misleading and result in unreasonable

suggestions for control limits.

The proposed methods are illustrated with the glass furnace production case,

which threads through the whole dissertation. However, the methods, especially the

Page 152: Topics in Multivariate Time Series Analysis: Statistical

134

dimension reduction procedure proposed in Chapter 4, has wider applications to many

business and economics fields.

6.1 Future Work

The proposed common cause chart and special cause chart largely depend on the

model fitted to the Phase I data. Undoubtedly, the estimation error from fitting the model

will add noises to the Phase II control charts. I would like to know by how much the

control charts are affected by estimation error. This research may answer the question of

how long Phase I should be run to collect enough data for constructing robust control

charts for Phase II.

The second branch of future work is to enhance the Box-Tiao method to deal with

more versatile time series models. The Box-Tiao method is based on assuming VARIMA

for the underlying process. How about ARCH, GARCH models? ARCH and GARCH

models have wide applications to economics and financial data. I would like to know

how the Box-Tiao method can be expanded to accommodate these models by applying

the same criteria – ranking components based on predictability.

In Chapter 4, I have discussed several spurious Box-Tiao components. The two-

step combined procedure can partly deal with some of them but not all, for example, the

components caused by lagged linear redundancy. For such problems, a linear cross-

sectional transformation falls short of a solution. It is thus of my interest to further look

into such cases.

Page 153: Topics in Multivariate Time Series Analysis: Statistical

135

APPENDIX A

EIGENVALUES INVARIANT TO TRANSFORMATION

For a 2-dimension VAR(1) model ttt azΦz += −1~~ , we can multiply each side by a

non-singular matrix A on the left,

1t t t−= +Az AΦz Aa% % , (A.1)

such that for the new innovation vector *t t=a Aa , we have ( )*

tCov =a I . If we transform

the observation vector as t t=y Az% %, then (A.1) can be written as,

1 * *1 1t t t t t t

−− −= = + = +y Az AΦA Az Aa Φ y a% % % % , (A.2)

where * 1−=Φ AΦA .

If iλ is an eigenvalue of Φ , i.e., i i iλ=Φe e , and we let i i=f Ae then,

( )* 1 1i i i i i i i i i iλ λ λ− −= = = = = =Φ f AΦA f AΦA Ae AΦe A e Ae f

Thus iλ is also an eigenvalue of *Φ . Without loss of generality, we can therefore

assume that 2 2a ×=Σ I , because for any covariance matrix we can always by a linear

transformation convert it to that form. Further, the eigenvalues, on which our discussion

is mainly based, are the same for Φand *Φ .

Page 154: Topics in Multivariate Time Series Analysis: Statistical

136

APPENDIX B

ARL CALCULATION FOR RESIDUAL BASED METHOD

If we assume that the time series model is correct and all parameters have been

estimated perfectly, then the residuals are independent. Under those assumptions the

calculation of the ARL for the residual based method is much simpler. Let N denotes the

number of steps after the step shift happens )( st = but before an observation falls outside

the control limits. Let 2tT denote the 2T statistics at time t, and c denotes the UCL.

( ) ( ) ( )1

2 2

0 0 1

n

t nn n t

ARL n P N n n P T c P T c−∞ ∞

= = =

= ⋅ = = ⋅ ≤ >

∑ ∑ ∏

If the process is under control and we have stpcTt >∀=> ,)Pr( 2 , then

1/ARL p= . However, for a step shift with a VAR(1) model that is no longer true.

Now consider the VAR(1) model ( )1 1t t t−= + − +z µ φ z µ a . Suppose we assume a

step change at time T given by: ( )t

t TE

t T

<=

+ ≥

µz

µ δ. Since we do not know when the

process mean has shifted, we estimate the residuals by ( ) ( )1t t t−= − − −e z µ Φ z µ . Thus

we have a residual mean shifted given by

( ) ( )( ) ( )( )

( )

1

0

1

t t tE E E

t T

t T

I t T

−= − − −

<

= = − = − ≥ +

e z µ Φ z µ

δ

δ Φδ Φ δ

Correspondingly, the non-centrality parameter induced by the shift is:

( )( ) ( )

( ) ( )

1 1

1

0

1

et t a t a

a

t T

t E E t T

I I t T

λ − −

< ′= ⋅Σ ⋅ = Σ = ′′ − Σ − ≥ +

e e δ δ

δ Φ Φ δ

Page 155: Topics in Multivariate Time Series Analysis: Statistical

137

Because of the dynamics of the time series model, a step shift in the mean of the

process causes two consecutive step shifts in the mean of the residuals. Therefore we

experience two consecutive step shifts of the non-centrality parameters of the mean of the

residuals. Typically, the first shift is upward and the second is downward, as illustrated in

Figure B.1. These shifts thus affects ( )2iP T c> accordingly. Now let ( )2

0 Tp P T c= >

denote the probability before the shift and 2( )tp P T c= > after the shift (t>T). We could

get a general formula for VAR(1) ARL:

( ) ( ) 2 00 0 02

1 11 1

1n

n

pARL p n p p p p p

p p

∞ −

=

−= + − − = − + −

Figure B.1. Non-centrality parameters shift.

Page 156: Topics in Multivariate Time Series Analysis: Statistical

138

APPENDIX C

PROOF OF (3.1)

By the definition of 1τ , i.e., starting from any in-control state, the number of steps

it takes to arrive at the out-of-control state for the first time, we have,

{ }

( ) { }1 0

0 0 01

min 1, s.t. 1| 1

| 1 min 1, s.t. 1| .

n

s

ni

n X s X s

P X i X s n X s X i

τ

=

= ≥ = + ≠ +

= = ≠ + ≥ = + =∑ (C.1)

Now let ( ) { }1 0min 1, s.t. 1|ni n X s X iτ = ≥ = + = , then (C.1) is

( ) ( )1 11

s

ii iτ φ τ

==∑ . Another relationship can be established by recognizing that starting

at state i , the MC can arrive at state 1s+ in the first step, or connect through state j at

the first step. Assuming ( ) ( )11|t tq i P X s X i−= = + = , we have

( ) ( ) ( )( )1 111 .

s

ijji q i r jτ τ

== + +∑ (C.2)

Now define ( ) ( ) ( )1 1 11 , ... , ...i sτ τ τ′ = τ and

( ) ( ) ( )1 , ... , ...q q i q s′ = q . By extending (C.2) we have

( )τ = q + R τ +1 . (C.3)

Thus ( ) ( ) ( )⋅ =-1 -1

τ = I - R q + R 1 I - R 1 . By substituting this back into (C.1) we

get ( )1τ ′=-1

φ I - R 1 . This is the same as (3.1).

Page 157: Topics in Multivariate Time Series Analysis: Statistical

139

APPENDIX D

DETERMING THE END STATES IN FIGURE 3-3

The ARMA(1, 1) model 1 1t t t tz z a aφ θ− −= + − can be written as:

( )1 1t t t ta z a zθ φ− −= + − .

This can be interpreted as a linear relationship between ta and tz where the slope

is 1 and the intercept is 1 1t t tb a zθ φ− −= − . If we let [ ]1 1 1t t tz a− − −′=w be the initial state

0S shown as the bold rectangle in Figure 3-3, the intercept tb will then range between

{ }1 0 1 1min min

tb S t ta zθ φ− ∈ − −= −w and { }

1 0 1 1max maxtb S t ta zθ φ− ∈ − −= −w . These two extremes

can only occur at the two different vertices of the bold rectangle. Thus with the initial

state fixed, the range of tb is fixed. Therefore the final states can only be those which

intersect with the lines between mint t ba z= + and maxt t ba z= + , as shown in Figure 3-3.

Page 158: Topics in Multivariate Time Series Analysis: Statistical

140

BIBLIOGRAPHY

Alt, F. B. (1984), “Multivariate quality control,” Encyclopedia of Statistical Sciences, eds. Kotz S., Johnson N. L. and Read C. R., New York: John Wiley, 110-122.

Alwan, A. J., and Alwan, L. C. (1994), “Monitoring Autocorrelated Processes Using Multivariate Quality Control Charts,” in Proceedings of the Decision Sciences Institute Annual Meeting, 3, 2106–2108.

Alwan, L. C., and Roberts, H. V. (1988), “Time-Series Modeling for Statistical Process Control,” Journal of Business & Economic Statistics, 6, 87-95.

Anderson, T. W. (1963), “The Use of Factor Analysis in the Statistical Analysis of Multiple Time Series,” Psychemetrika 28, 1-25.

Richard G. A., Qian, A., and Rasche, R. H., “Analysis of Panel Vector Error Correction Models Using Maximum Likelihood, the Bootstrap, and Canonical-Correlation Estimators,” unpublished manuscript, August 2006.

Apley, D. W., and Tsung, F. (2002), “The Autoregressive T2 Chart for Monitoring Univariate Autocorrelated Processes,” Journal of Quality Technology, 34, 80-96.

Atienza, O. O., Tang, L. C., and Ang, B. W. (2002), “A Cusum Scheme for Autocorrelated Observations,” Journal of Quality Technology, 34, 187-199.

Badcock, J., Vailey, T. C., Jonathan, P., and Krzanowski, W. J. (2004), “Two Projection Methods for Use in the Analysis of Multivariate Process Data With an Illustration in Petrochemical Production,” Technometrics. 46, 392-403.

Bagshaw, M., and Johnson, R. A. (1975), “The Effect of Serial Correlation on the Performance of CUSUM Tests II,” Technometrics, 17, 73-80.

Berthouex, P. M., Hunter, W. G., and Pallesen, L. (1978), “Monitoring Sewage Treatment Plants: Some Quality Control Aspects,” Journal of Quality Control, 10, 139-149.

Bewley, R., Orden, D., Yang, M., and Fisher, L. A. (1994), “Comparison of Box-Tiao and Johansen canonical estimators of integrating vectors in VEC(1) models,” Journal of Econometrics, 64, 3-27.

Bewley, R., and Yang, M. (1995), “Tests for cointegration based on canonical correlation analysis,” Journal of the American Statistical Association, 90, 990-996.

Bisgaard, S., and Huang, X. (2008), “Visualizing Principal Components Analysis for Multivariate Process Data,” Journal of Quality Technology, 40, 334-344.

Black, F., Jensen, M. C., and Scholes, M., 1972, “The Capital Pricing Model: Some Empirical Testing,” Studies in the Theory of Capital Market, Jensen, M. C., Ed., New York: Praeger Publisher.

Page 159: Topics in Multivariate Time Series Analysis: Statistical

141

Brockwell, P. J., and Davis, R. A. (2002), Introduction to Time Series and Forecasting (2nd ed.), New York: Springer-Verlag.

Bossaerts, P. (1988), “Common nonstationary components of asset prices,” Journal of Economic Dynamics and Control, 12, 347-364.

Brook, D., and Evan, D. A. (1972), “An Approach to the Probability Distribution of CUSUM RL,” Biometrika, 59, 539-549.

Box, G. E. P., and Jenkins, G. M. (1968), “Some recent advances in forecasting and control. Part 1,” Applied Statistics, 17, 91-109.

Box, G. E. P., Jenkins, G. M., and MacGregor, J. F. (1974), “Some Recent Advances in Forecasting and Control,” Applied Statistics, 23, 158-179.

Box, G. E. P., and Jenkins, G. M. (1976), Time Series Analysis: Forecasting and Control (2nd ed.), San Francisco, CA: Holden-Day Inc.

Box, G. E. P., and Tiao, G. C. (1977), “A Canonical Analysis of Multiple Time Series,” Biometrika 64, 355-365.

Box, G. E. P. (1991), “Bounded Adjustment Charts,” Quality Engineering, 4, 331-338.

Box, G. E. P., Jenkins, G. M., and Reinsel, G. C. (1994), Time Series Analysis: Forecasting and Control (3rd ed.), New Jersey: Prentice-Hall, Inc.

Box, G. E. P., and Luceño, A. (1997), Statistical Control by Monitoring and Feedback Adjustment, New York: John & Wiley & Sons, Inc.

Box, G. E. P., Hunter, J. S., and Hunter, W. G. (2005), Statistics for Experimenters: Design, Innovation, and Discovery (2nd ed.), New York: Wiley-Interscience.

Box, G. E. P., and Paniagua-Quiñones, C. (2007), “Two Charts: Not One,” Quality Engineering, 19, 93-100.

Brockwell, P. J., and Davis, R. A. (2002), Introduction to Time Series and Forecasting, New York: Springer.

Campbell, J. Y., et al., 2000, “Have Individual Stocks Become More Volatile? An Empirical Exploration of Idiosyncratic Risk,” Journal of Finance, 56, 1-43.

Chamberlain, G., and Rothschild, M. (1983), “Arbitrage, Factor Structure, and Mean-Variance Analysis on Large Asset Markets,” Econometrica, 51, 1281-1304.

Chen, N., Roll, R. and Ross, S. (1986), “Economic Forces and Stock Market,” Journal of Business, 59, 383-403.

Cleveland, W. S. (1985), The Elements of Graphing Data, North Scituate, MA: Duxbury Press.

Page 160: Topics in Multivariate Time Series Analysis: Statistical

142

Cleveland, W. S. (1993), Visualizing Data, Summit, NJ: Hobart Press.

Connor, G., and Korajczyk, R. (1987), “Risk and Return in an Equilibrium APT,” Journal of Financial Economics, 21, 255-289.

Cressie, N. A. C. (1993), Statistics for Spatial Data, New York: John Wiley & Sons.

Crosier, R. B. (1988), “Multivariate Generalizations of CUSUM Quality-Control Schemes,” Technometrics, 30, 291-303.

Deming, W. E. (1982), Quality, Productivity and Competitive Position, Cambridge, MA: MIT Center for Advanced Engineering Study.

Dickey, D. A. and Fuller, W. A. (1979), “Distribution of the Estimators for Autoregressive Time Series with a Unit Root,” Journal of the American Statistical Association, 74, 427-431.

Durbin, J., Koopman, S. J. (2001), Time Series Analysis by State Space Method, Oxford University Press.

Engle, R. F., and Granger, C. W. J. (1987), “Co-integration and Error Correction Representation, Estimation and Testing,” Econometrica, 55, 251-276.

Engle, R. F., and Yoo, S. (1987), “Forecasting and Testing in Cointegrated Systems,” Journal of Econometrics, 35, 143-59.

Fama, E. F. and French, K. R. (1992), “The Cross-Section of Expected Stock Returns,” The Journal of Finance, 2, 427-465.

Fama, E. F. and French, K. R. (1993), “Common Risk Factors in the Returns on Stock and Bonds,” Journal of Financial Economics, 33, 3-56.

Gabriel, K. R. (1985), “Multivariate Graphics,” Encyclopedia of Statistical Sciences, eds. D. L. Banks, C. B. Read and S. Kotz, New York: John Wiley and Sons., Vol. 6, 66-79.

Gnanadesikan, R. (1997), Methods for Statistical Data Analysis of Multivariate Observations, New York: John Wiley & Sons.

Hald, A. (1951), Statistics with Engineering Applications, New York: John Wiley and Sons.

Harris, T. J. and Ross, W. H. (1991), “Statistical Process Control Procedures for Correlated Observations,” The Canadian Journal of Chemical Engineering, 69, 48-57.

Page 161: Topics in Multivariate Time Series Analysis: Statistical

143

Harvey, A. C., Todd, P. H. J. (1983), “Forecasting Economic Time Series with Structural and Box-Jenkins Models: A Case Study,” Journal of Business & Economic Statistics, 1, 299-307.

Holmes, D.S. and Mergen, A.E. (1993), “Improving the Performance of The T2 Control Chart,” Quality Engineering, 5, 619-625.

Hotelling, H. (1947), “Multivariate Quality Control – Illustrated by the Air Testing of Sample Bombsights,” Techniques f Statistical Analysis, eds. C. Eisenhart, M. W. Hastay, and W. A. Wallis, New York: McGraw-Hill, 111-184.

Huang, X., and Bisgaard, S. (2009), “A Class of Markov Chain Models for Average Run Length Computations for Autocorrelated Processes,” Technometrics, tentatively accepted.

------.(2010), “Model-Based Multivariate Monitoring Charts for Autocorrelated Processes,” submitted.

Jackson, J. E. (1985), “Multivariate Quality Control,” Communications in Statistics – Theory and Methods, 14, 2657-2688.

Jackson, J. E. (1991), A User's Guide to Principal Components, New York: John Wiley & Sons.

Jiang, W. (2001), “Average Run Length Computation of ARMA Charts for Stationary Process,” Commun. Statist. – Simula., 30, 699-716.

Jiang, W. (2004), “Multivariate Control Charts for Monitoring Autocorrelated Processes,” Journal of Quality Technology, 36, 367-379.

Jiang, W., Tsui, K. L., and Woodall, W. H. (2000), “A New SPC Monitoring Method: The ARMA Chart,” Technometrics, 42, 399-410.

Johansen, S. (1988), “Statistical Analysis of Co-integration Vectors,” Journal of Economic Dynamic and Control, 12, 231-254.

Johnson, R. A., and Bagshaw, M. (1974), “The Effect of Serial Correlation on the Performance of CUSUM Tests,” Technometrics, 16, 103-112.

Johnson, R. A., and Langeland, T. (1991), “A Linear Combinations Test for Detecting Serial Correlation in Multivariate Samples,” Topics in Statistical Dependence, eds. Block, H. et al., Institute of Mathematical Statistics Monograph, 299-313.

Johnson, R. A., and Wichern, D. W. (2002), Applied Multivariate Statistical Analysis (5th ed.), New York: Prentice Hall.

Jolliffe, I. T. (2002), Principal component analysis, New York: Springer.

Page 162: Topics in Multivariate Time Series Analysis: Statistical

144

Juran, J. M. (1986), “The Quality Trilogy: A Universal Approach to Managing for Quality,” Quality Progress, August, 19-24.

Kourti, T., and MacGregor, J. F. (1996), “Multivariate SPC Methods for Process and Product Monitoring,” Journal of Quality Technology, 28, 409-428.

Kramer, H., and Schmid, W. (1997), “EWMA Charts for Multivariate Time Series,” Sequential Analysis, 16, 131-154.

Krieger, C. A., Champ, C. W., and Alwan, L. C. (1992), “Monitoring an Autocorrelated Process,” presented at the Pittsburgh Conference on Modeling and Simulation, Pittsburgh, PA.

Lehmann, B. N. and Modest, D. (1988), “The Empirical Foundations of the arbitrage pricing theory,” Journal of Financial Economics, 21, 213-254.

Lin, W. S., and Adams, B. M. (1996), “Combined Control Charts for Forecast-based Monitoring Schemes,” Journal of Quality Technology, 28, 289-301.

Lintner, J. (1965), “The Valuation of Risk Assets and the Selection of Risky Investments in Stock Portfolios and Capital Budgets,” Review of Economics and Statistics, 47, 13-37.

Lintner, J. (1969), “The Aggregation of Investor’s Diverse Judgments and Preference in Purely Competitive Security Markets,” Journal of Finance and Quantitative Analysis, 4, 347-400.

Loredo, E. N., Jearkpaporn, D., and Borror, C. M. (2002), “Model-based Control Chart for Autoregressive and Correlated Data,” Quality and Reliability Engineering International, 18, 489-496.

Lowry, C. A., Woodall, W. H., Champ, C. W., Rigdon, S. E.. (1992), “A Multivariate Exponentially Weighted Moving Average Control Chart,” Technometrics, 34, 46-53.

Lowry, C.A., and Montgomery, D.C. (1995), “A Review of Multivariate Control Charts,” IIE Transactions, 27, 800-810.

Lu, C. W., and Reynolds, M. R., Jr. (1999a), “Control Charts for Monitoring the Mean and Variance of Autocorrelated Processes,” Journal of Quality Technology, 31, 259-274.

------.(1999b), “EWMA Control Charts for Monitoring the Mean of Autocorrelated Processes,” Journal of Quality Technology, 31, 166-188.

------.(2001), “Cusum Charts for Monitoring an Autocorrelated Processes,” Journal of Quality Technology, 33, 316-334.

Page 163: Topics in Multivariate Time Series Analysis: Statistical

145

Lucas, J. M., Saccucci, M. S. (1990), “Exponentially Weighted Moving Average Control Schemes: Properties and Enhancements,” Technometrics, 32, 1-12.

Lütkepohl, H. (1993), Introduction to Multiple Time Series Analysis, New York: Springer-Verlag.

Mason, R. L., and Young, J. C. (2002), Multivariate Statistical Process Control with Industrial Applications, Philadelphia: ASA-SIAM.

Mastrangelo, C. M., Runger, G. C., and Montgomery, D. C. (1998), “Statistical Process Monitoring with Principal Components,” Quality and Reliability Engineering International, 12, 203-210.

Merton, R. C. (1973), “An intertemporal capital asset pricing model,” Econometrica, 41, 867-887.

Molenaar, P. C. M. (1985), “A Dynamic Factor Model for the Analysis of Multivariate Time Series,” Psychometrika. 50, 181-202.

Montgomery, D. C., and Mastrangelo, C. M. (1991), “Some Statistical Process Control Methods for Autocorrelated Data,” Journal of Quality Technology, 23, 179-193.

Montgomery, D. C. (2008), Introduction to Statistical Quality Control (6th ed.), Wiley.

Mossin, J. (1966), “Equilibrium in a Capital Asset Market,” Econometrica, 768-783.

Pan, X. (2005), “Notes on Shift Effects For T2-type Charts on Multivariate ARMA Residuals,” Computers & Industrial Engineering, 49, 381-392.

Peña, D., and Box, G. E. P. (1987), “Identifying a Simplifying Structure in Time Series,” Journal of the American Statistical Association, 82, 836-843.

Peña, D., and Poncela, P. (2004), “Forecasting with nonstationary dynamic factor models,” Journal of Econometrics, 119, 291-321.

Peña, D., Tiao, G. C., and Tsay, R. S. (2001), A Course in Time Series Analysis, New York: John Wiley & Sons.

Pignatiello, J. J. and Runger, G. C. (1990), “Comparisons of multivariate CUSUM charts,” Journal of Quality Technology, 22, 173-186.

Playfair, W. (1786), The Commercial and Political Atlas, Cambridge: Cambridge University Press.

Quenouille, M. H. (1957), The Analysis of Multiple Time Series, New York: Hafner.

Reinsel, G. (1983), “Some Results on Multivariate Autoregressive Index Models,” Biometrika. 70, 145-156.

Page 164: Topics in Multivariate Time Series Analysis: Statistical

146

Reinsel, G. C. (1997), Elements of Multivariate Time Series Analysis, New York: Springer.

Rencher, A. C. (2002), Methods of Multivariate Analysis, New York: John Wiley & Sons.

Roll, R. and Ross, S. A. (1980), “An Empirical Investigation of the Arbitrage Pricing Theory,” The Journal of Finance, 35, 1073-1103.

Rosolowski, M. and Schmid, W. (2003), “EWMA Charts for Monitoring the Mean and Autocovariances of Stationary Processes,” Statistical Papers, 47, 595-630.

Ross, S. A. (1976), The Arbitrage Theory of Capital Asset Pricing, Journal of Economic Theory, 13, 341-360.

Runger, G. C. and Alt, F. B. (1996), “Choosing principal components for multivariate statistical process control,” Communications in Statistics: Theory and Methods, 25, 909-922.

Runger, G. C. and Prabhu, S. S. (1996), “A Markov Chain Model for the Multivariate Exponentially Weighted Moving Averages Control Chart,” Journal of American Statistical Association, 91, 1701-1706.

Runger, G. C., Willemain, T. R. and Prabhu, S. S. (1995), “Average Run Lengths for Cusum Control Charts Applied to Residuals,” Communications in Statistics - Theory and Methods, 24, 273-282.

Seber (2004), Multivariate Observations, New York: John Wiley & Sons.

Sharpe, W. F. (1963), “A Simplified Model for Portfolio Analysis,” Management Science, 9, 277-293.

Sharpe, W. F. (1964), “Capital Asset Prices: A Theory of Market Equilibrium under Conditions of Risk,” Journal of Finance, 19, 425-442.

Shewhart, W. A. (1931), Economic Control of Quality of Manufactured Product, New York: Van Nostrand.

Shu, L. J., Apley, D. W., and Tsung, F. (2002), “Autocorrelated Processes Monitoring Using Triggered Cuscore Charts,” Quality and Reliability Engineering International, 18, 411-421.

Sloan, I. H., and Joe, S. (1994), Lattice Methods for Multiple Integration, Oxford: Clarendon Press.

Sullivan, J. H., and Woodall, W. H. (1996), “A Comparison of Multivariate Control Charts for Individual Observations,” Journal of Quality Technology, 28, 398-408.

Page 165: Topics in Multivariate Time Series Analysis: Statistical

147

Superville, C. R., and Adams, B. M. (1994), “An Evaluation of Forecast-based Quality Control Schemes,” Communications in Statistics - Simulation and Computation, 23, 645-661.

Switzer, P. (1985), “Min/Max Autocorrelation Factors for Multivariate Spatial Imagery,” in Computer Science and Statistics: Proceedings of the 16th Symposium on the Interface, 13-16.

Tiao, G. C., and Box, G. E. P. (1981), “Modeling multiple time series with applications,” Journal of the American Statistical Association, 76, 802-816.

Tiao, G. C., and Tsay, R. S. (1989), “Model Specification in Multivariate Time Series,” Journal of Royal Statistical Society. 51, 157-213.

Tiao, G. C., Tsay, R. S., and Wang, T. (1993), “Usefulness of linear transformations in multivariate time-series analysis,” Empirical Economics 18, 567-593.

Tracy, C. D., Young, J. C., and Mason, R. L. (1992), “Multivariate Control Charts for Individual Observations,” Journal of Quality Technology, 24, 88-95.

Treynor, J. (1961), “Toward a Theory of the Market Value of Risky Assets,” unpublished manuscript.

Tsay, R. S. (2005), Analysis of Financial Time Series (2nd ed.), New York: John Wiley and Sons.

Tufte, E. R. (1983), The Visual Display of Quantitative Information, Cheshire: Graphic Press.

Tukey, J. W. (1962), “The Future of Data Analysis,” Annals of Mathematical Statistics 33, 1-67.

Vanbrackle, L. N., and Reynolds M. R. (1997), “EWMA and Cusum Control Charts in the Presence of Correlation,” Communications in Statistics - Simulation and Computation, 26, 979-1008.

Velu, R. P., Reinsel, G. C., and Wichern, D. W. (1986), “Reduced Rank Models for Multiple Time Series,” Biometrika, 73, 105-118.

Velu, R. P., Wichern, D. W. and Reinsel, G. C. (1987), “A note on non-stationarity and canonical analysis of multiple time series models,” Journal of Time Series Analysis, 8, 479-487.

Wainer, H. (2004), Graphic Discovery: A Trout in the Milk and Other Visual Adventure, Princeton: Princeton University Press.

Wardell, D.G., Moskowitz, H. and Plante, R.D. (1992), “Control Charts in the Presence of Data Correlation,” Management Sciences, 38, 1084-1105.

Page 166: Topics in Multivariate Time Series Analysis: Statistical

148

Wardell, D.G., Moskowitz, H., and Plante, R.D. (1994), “Run-Length Distribution of Special-Cause Control Charts for Correlated Processes,” Techonometrics, 36, 3-17.

Wierda, S. J. (1994), “Multivariate statistical process control - Recent results and directions for future research,” Statistica Neerlandica, 48, 147-168.

Woodall, W. H., and Ncube, M. M. (1985), “Multivariate CUSUM quality-control procedures,” Technometrics, 27, 285-292.

Woodall, W. H., (2000), “Controversies and Contradictions in Statistical Process Control,” Journal of Quality Technology, 32, 341-350.